You are on page 1of 57

IDENTIFICATION OF

MINERALS
IN
HAND SPECIMEN
AND
THIN SECTION

Andrew G. Christy,

Research School of Earth Sciences,


Australian National University

4th edition, February 2010


INTRODUCTION

A mineral species is a naturally occurring chemical compound that may contain


various solid solution components but is dominated by one specified such end-member, and
which occurs with a particular crystal structure type. For instance, a carbonate which is
predominantly CaCO3 in a particular structure with trigonal-rhombohedral symmetry is
called calcite. Minerals MCO3 with the same structure in which other metals predominate
over calcite have different names (M = Fe: siderite; M = Mg: magnesite, and so on). The
mineral CaCO3 with a different structure (stable at higher pressures) is a different species
called aragonite. Although more than 4000 species have now been described, very few of
these are important as rock-forming minerals or in ore deposits. The vast majority of rock-
forming minerals are compounds of only the ten or so most abundant chemical elements
(Fig. 1). In particular, 40% of mineral species are silicates, containing as essential constituent
Si and O, the two most abundant elements in the crust! Because mineral compounds have a
wide range of possible cations (positively charged) but a much smaller variety of anions
(negatively charged) in their make-up, minerals are frequently classified according to the
anion, into groups such as oxides, sulphides, carbonates and silicates. The element symbols
of Fig.1 are defined in Table C1, Appendix C. Numerical abundances are in Table C5, along
with primitive Solar System, mantle and seawater data.

Some rarer elements are readily concentrated by geochemical processes to produce a


disproportionate diversity of mineral species; examples are the compounds of the
economically important base metals Cu, Zn and Pb (look at these in Fig. 2). The chemistry
of minerals that you are likely to meet will place them in either the “rock-forming mineral”
group or this one. Both the common and rarer chemical elements are likely to form only
particular types of chemical compounds as a result of their differing chemical preferences.
This can be expressed using a version of Goldschmidt’s “siderophile” (metal-loving) -
“chalcophile” (sulphide-loving) - “lithophile” (oxide/silicate -loving) classification (Fig. 3).
Knowledge of this behaviour can help with identification: for instance, a known sulphide is
unlikely to contain beryllium or uranium! This data can also be seen in Table C1.

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 1


Note for chemists: the “philicity” of an element can be quantified as a function of the free
energies of formation of the oxide and sulphide, and is related to both its electronegativity
and its Pearson hardness/softness.

Identification is further helped by the fact that many of the species are naturally
grouped into large, distinctive families whose members are much more similar in crystal
structure and some physical properties to each other than they are to non-members (e.g.
pyroxenes, garnets, amphiboles). These groups can be easy to identify. The individual species
within them are distinguished mainly on the basis of chemical composition, and may have
distinctive regions of pressure-temperature stability (blue amphiboles in blueschists, green
amphiboles in greenschists, black amphiboles in amphibolites!) but colour and associated
minerals provide very useful clues. Familiarity with a few tens of minerals or mineral groups
will allow you to identify most major minerals in most rocks rapidly, and allow you to home
in on the more obscure ones efficiently.

Mineral species are defined rigorously according to chemical composition and crystal
structure, which these days are usually determined by professional researchers using
techniques such as electron microprobe analysis and X-ray diffraction. Unlike pure end-
member chemical compounds, minerals show considerable variation in chemical
composition, which can only be understood in the context of some basic knowledge of their
crystal structure, and which contains a wealth of information about the conditions under
which they formed. However, the descriptive groundwork for the major rock-forming
minerals and ores was in place a long time before X-rays or electron beams were available, or
the subtleties of solid solutions and phase equilibria had been investigated. The 18th century
pioneers made ingenious use of physical properties to identify minerals, especially optical
properties as seen in a specially adapted microscope using linearly polarised light. The
sample preparation and apparatus involved are relatively simple and cheap, so we still use
these techniques today for initial identification and fast reconnaissance of rocks and minerals.

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 2


This handbook is about how to use these techniques, and what information can be gained
from them.

Much mineral identification can be done by learning:

(i) what features to look for in a specimen,


(ii) what not to get distracted by,
(iii) what each direct observation is telling you about other features that you cannot see
directly, such as the composition and crystal structure.

A few important approaches are outlined below.

PETROLOGICAL CONTEXT

If several minerals are in a rock, forming an assemblage of other species, then


knowledge of which species commonly associate with each other and which do not (ie,
petrology), means that identfying one or two can be a great help with identifying the rest.
Table C4 in Appendix C groups minerals by the types of rock in which they occur.

The first question to ask is: Is this rock sedimentary, metamorphic or igneous?

Minerals formed during diagenesis of a sedimentary rock will necessarily be those stable at
low-temperatures: clays, carbonates etc. There may be clasts that are fragments of higher-
grade rocks, and there may be biogenic minerals such as carbonates and phosphates.

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 3


Metamorphic rocks will hopefully show a peak assemblage of minerals which crystallised
together as relatively large crystals at the highest temperature experienced. Not all minerals
are necessarily simultaneous in their growth, though. Metamorphic rocks usually have
textures that record deformation events, and that the growth of some minerals may be timed
differently relative to particular events. There may also be finer-grained retrogressive
minerals formed by alteration on the way back down to room temperature. The peak
assemblage will depend on the bulk composition of the rock, and the pressure and
temperature reached. Different minerals have different stability limits. For example, you will
not find kyanite in a rock that equilibrated at low pressure or very high temperature
Andalusite is the stable aluminosilicate at low pressure and moderate temperatures, while
sillimanite replaces both of these in high temperature rocks. Identifying one or two minerals
may be enough to pin down a pressure-temperature range, or metamorphic facies.

Igneous rocks crystallise from a melt, and hence show a sequential fabric in which first one
mineral forms crystals, then a second, then a third, and so on. The exact order is very
sensitive to bulk composition and pressure of crystallisation. The vast majority of igneous
rocks are made up of the minerals of Bowen’s sequences: a Mg-Fe rich silicate or two which
become increasingly polymerised as rocks differentiate to more Si-rich compositions (olivine
> pyroxene > amphibole > dark mica) and Mg-Fe poor silicates which are usually Ca-rich
plagioclase in less differentiated (more mafic) rocks and Na-rich plagioclase, K-feldspar and
quartz in more felsic rocks. Alkaline rocks will have feldspathoids (eg nepheline, analcime,
leucite) instead of quartz. No rock will have these AND quartz since the combination is not
stable! Many igenous rocks contain low-temperature alteration products. Make sure that you
can tell these from the fresh high-temperature minerals.

(Chalcophile) Ores do not always fit neatly into the categories since they usually have been
formed by unusual processes, and have a highly abnormal bulk composition! “Incompatible”
elements like Cu, Pb and Zn do not go readily into the normal rock-forming minerals, and
form segregations of their own distinctive minerals. Often, they concentrate in liquids which
form veins cutting through other rocks, and may react with those rocks, causing alteration.
Primary ores of chalcophile elements are sulphides. In the oxidised zone of a deposit, these
are replaced by oxides, carbonates, sulphates etc. Ore-bearing fluids often contain
incompatible lithophile elements that crystallise as “gangue minerals” such as fluorite
(containing F) and barite (containing Ba).

Other questions that can help you use one mineral to identify others:

• What does your identified species imply about bulk composition, rock type or
protolith, and major elements present?

• Does it eliminate any incompatible minerals, or suggest other minerals which are
likely to be associated with your known species?

• Do your deductions about composition and pressure-temperature range eliminate or


support particular species names for pyroxene, amphibole, garnet etc (eg pyrope
garnet only stable in eclogite facies, but almandine stable at lower pressure)?

PHYSICAL AND CHEMICAL PROPERTY TESTS

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 4


If you have a hand specimen available, “looking” may mean not just visual
identification but also:

(i) hefting it in your hand: you rapidly develop a “feel” for specimens which are more than
about 25% denser than the crustal average, and which are likely to contain heavy (high
atomic number) elements such as barium (Ba) or lead (Pb).

(ii) breaking off small crumbs and checking for magnetism. Magnetite (Fe3O4) is the only
common strongly magnetic mineral. Other Fe minerals such as ilmenite and pyrrhotite also
show a weaker response, however. Many other black or brassy/bronzy minerals do not.

(iii) checking a small crumb for water solubility. Cubes of halite (NaCl) and sylvite (KCl)
dissolve, while cubes of fluorite do not. Most other soluble minerals are relatively rare
sulphates, also found in evaporite deposits and as mine/cave/quarry efflorescences. Tasting
minerals is no longer considered advisable, since some soluble species are quire poisonous.

(iv) a drop of hydrochloric acid (about 10% strength) will identify carbonates by their fizzing
(carbon dioxide released), and may distinguish calcite (fast reaction) from dolomite (slow
reaction).

(v) some species can produce distinctive smells. Arsenic-bearing sulphides frequently give of
a garlic-like smell of arsine when struck with a hammer. Many soluble sulphates have an
acrid reek of sulphuric acid. The more soluble sulphides liberate eggy hydrogen sulphide if
reacted with acids.

DETAILED MINERALOGICAL EXAMINATION

First look at the form taken by your unknown mineral.


• Does it occur in crystals of a distinctive shape (eg “needles”)?
• Do the crystals aggregate to form larger-scale distinctive shapes (eg “balls of
radiating needles”)?
• Do broken surfaces show you cleavage or fracture?
• Is a small, broken-off crumb brittle, or does it show distinctive tenacity behaviour?
• What is the Mohs hardness?

These tests all give strong clues as to the general chemical type of the mineral, and its crystal
symmetry.

Second, look at the optical properties. There is much more to these than just “colour”.
The colour of a hand specimen, the colour of a fine powder (streak), and the various colours
that can be seen in a grain mount or thin section under plane-polarised and cross-polarised
light all complement each other to give a wealth of information about the material.

These observation techniques are discussed in more detail below. Many of the
properties discussed provide clues about the type of chemical bonding and the crystal
structure of the mineral. A summary of basic crystallographic concepts and terms is provided
in Appendix A.

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 5


HABIT

Before any more detailed testing, you should look at the overall form taken by a
mineral. The word “habit” is used to mean three somewhat different things:

(i) the exact combination of crystallographic forms that make up the faces of a crystal,
(ii) the overall shape of a single crystal
(iii) the form taken by aggregates of crystals.

A given mineral often has a limited range of habits. Examples of habit include:

Crystal Forms (see Appendix A, Fig. A6):


Cube: cubic {100} faces
Octahedron: {111} in three of the five cubic crystal classes
Tetrahedron: {111} or {-111} in the other two cubic classes.
Rhombic Dodecahedron: cubic {110} faces; ideally twelve rhombs.
Trapezohedron: a twenty-four faced cubic form such as {211}
Pyramid: symmetry-related faces that converge at a point on
a high-symmetry axis.
Dipyramid: two sets of pyramid faces back-to-back, related by a
horizontal mirror plane.
Prism: a set of symmetry-related faces that are all parallel to the
same high-symmetry axis
Pinacoid: a pair of opposite faces that are related to no others by
symmetry. Most often the “basal pinacoid” or {001} faces
(normal to the high-symmetry axis) of a trigonal, tetragonal
or hexagonal crystal.

These are the most common survivors out of a huge vocabulary of 19th century
jargon, which is gradually dying out!

Habit of Single Crystals:


Equant: Roughly the same dimensions in all three directions. Undistorted crystals
of the cubic system are all equant.
Prismatic: Long and thin due to well-developed prism faces
Acicular: Needle-like: very long and thin.
Fibrous: Extremely long and thin!
Tabular: Platy. Short and fat due to well-developed pinacoids (see above).
Flaky, Scaly: Extremely tabular.
Blady: One long direction, one thin direction, and one that is in between, like a
knife blade.

The habit of crystals can be an important clue to the likely crystal symmetry.
However, because growth can be distorted by asymmetries in the environment such as
presence of neighbouring crystals, fluid currents and temperature gradients , it is not reliable.
Fortunately, angles between faces are not affected by changes in the relative size of faces,
hence the usefulness of goniometry!

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 6


Striations:
Some minerals oscillate between two different crystal forms as they grow. This produces not-
quite-flat faces with “stripes” or ledges across them. The pattern of striations can be
characteristic for a particular mineral. Quartz prisms often show striations across the width
of the prism faces, while tourmaline striations are along the length of the prism. Pyrite
cubes often show striations demonstrating that the real rotation symmetry of the cube face is
only 2-fold, not 4-fold.

Habit of Aggregates
Botryoidal: Like a bunch of grapes: partially merged balls, which are often made of
concentric shells or radiating fibres.
Columnar: Semi-parallel prisms
Reticulated: “Jackstraw”. Interlocked, intergrown prisms with gaps running between
them.
Boxwork: Honeycomb-like shells surrounding voids. The mineral probably grew in
fractures or cleavages in an earlier matrix phase, which has since dissolved
away.
Arborescent: Tree-like, like some native copper.
Dendritic: Delicate moss-like or tree-like, like some Mn oxide precipitates.

TWINNING

Twinned crystals have part of the crystal structure “on backwards” relative to the rest.
Like cleavage (below) the orientation of twin components is strongly controlled by the crystal
structure. Only a few minerals show twinning commonly on the naked-eye scale. Watch for
the cross-shaped twins of staurolite, elbow-shaped twins of rutile, simple “Carlsbad” twins
of orthoclase feldspar in which the two interpenetrating halves show different lustres and
surface texture, and stripy lamellar twinning in plagioclase. Twins are discussed further in the
section on microscopy below.

CLEAVAGE AND FRACTURE

The way in which a mineral breaks offers several clues about its crystal structure. It is
not always necessary to damage a specimen in order to determine cleavage or fracture: most
mineral specimens already have a few fresh surfaces exposed.

Cleavages
Are the very flat, mirror-like surfaces produced by breakage on a plane of weak
bonding in the crystal structure. A mineral can be split easily on a plane of good cleavage
anywhere through its bulk. Cleavage is often distinguished as “perfect”, “good”, “fair”,
“poor/weak”, etc. Only perfect-to-good cleavages are easy to recognise.
A cleavage plane will display the same quality of cleavage as every other plane that is
related to it by symmetry. Therefore, it is impossible for a cubic crystal to show less than
three different orientations of cleavage plane. Note that these can look quite different
depending on your viewing direction (Fig. 4).

Cubic possibilities are:


Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 7
Cubic cleavage: 3 cleavages all at 90o, parallel to {100} faces. Galena, halite.
Octahedral cleavage: planes in 4 orientations parallel to {111}. Fluorite. Note that
many fluorite crystals have a cube-shaped growth form,
but the corners break off easily to give octahedral cleavage
forms.
Rhombic dodecahedral: planes in 6 orientations parallel to {110}. Sphalerite,
sodalite.

Non-cubic possibilities include:

Basal: One unique cleavage, usually across a prism direction. Micas and other
layer silicates like chlorites, smectites, apophyllite. These all have a perfect
basal cleavage and split into very thin flakes and sheets easily.
The prismatic crystals of apatite and beryl can show a poor basal
cleavage. Prisms of topaz have a perfect basal cleavage, which gives broken
crystals their very flat, rather perfect bottom.

Prismatic: one or more orientations running along the length of a crystal. Examples
include:

1 only:
Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 8
Sillimanite forms “square” prisms with only one good cleavage parallel to one
diagonal of the “square”. It is clearly lower in symmetry than tetragonal!
Epidote minerals have one perfect cleavge {001} parallel to the usual long
directiony.

2, unrelated by symmetry:
Feldspars have two cleavages {001} and {010} that are nearly at
right-angles but are not related by symmetry to each other (two different
pinacoidal cleavages, one slightly more perfect than the other!).
Kyanite has two cleavages of very different quality at about 74o:
perfect {100} parallel to the blades and good {001} along their length, as
well as a frequent {001} parting (see below) across their length (Fig. 5).

2, related by symmetry:
Very common in chain silicates. Pyroxenes have two equivalent cleavages at
nearly 90o, amphiboles have two at nearly 120o. Hence, pyroxenes have a
“blocky” look on broken surfaces, whereas amphiboles have a splintery, flaky
look.

3, unrelated by symmetry
The pyroxenoid wollastonite has three cleavages in the same zone {h0l},
making it very splintery.

Barite has two symmetry-related {210} cleavages at 78o, and an unrelated


{001} basal cleavage perpendicular to these two. Be careful to distinguish
these from the three equivalent rhombohedral cleavages of calcite, where each
pair of planes makes traces at an angle of 78o on the third.

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 9


3, related by symmetry
Hexagonal prismatic perfect in the feldspathoid cancrinite, poor in nepheline.

4, in two-symmetry related pairs


Scapolite has prismatic cleavages {100} and {110} at 45o, making it rather
splintery.

Rhombohedral: The distinctive {10-14} cleavage of the calcite and dolomite group
carbonates. Three equivalent planes all at about 100o to each other.

Partings
These can look like a cleavage but are due to the presence of twin planes, precipitates
or inclusions in thin layers (lamellae). The mineral only splits here and there, and not all
pieces will display the parting. Basal and rhombohedral parting in hematite , ilmenite and
corundum is an example.

Fracture
The type of surface obtained from a material with poor or non-existent cleavages.
Terms that are used include:

Conchoidal: shell-shaped flakes, like broken glass. Characteristic of crystals with no


cleavage that
are strongly bonded in all directions, like quartz, olivine and garnet.

Uneven: the texture of fine-granular aggregates.

Hackly: pulled into sharp points, as happens with ductile native metals.

Splintery: self-explanatory: fibrous aggregates.

TENACITY

This is another mechanical property of the material which may involve destroying a
small piece of a sample! Tenacity encapsulates the ease and mechanisms by which a mineral
deforms. A mineral may be:

Brittle: ie, it breaks into small pieces if you hit it hard enough. Most minerals are
brittle.

Sectile: the mineral is just “plastic” enough to be able to whittle small turnings off
with a knife. Many softer minerals of heavy metals like Ag, Sb, Hg and Pb.
Acanthite, chlorargyrite, stibnite, calomel.

Flexible: strongly bonded in at least one direction, but loosely enough bonded in
other directions that a crystal can “flow”, and can be bent without breaking.

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 10


Stibnite, molybdenite, micas, chlorite.

Elastic: A crystal springs back after being bent (within limits): mica,
but not chlorite.

Malleable: Samples can be flattened, stretched, smeared, rolled or bent into a variety of
shapes. Ductility is the ability to be drawn out into a thin wire, which goes
hand-in-hand. Most native metals such as copper, silver, gold and lead are
very malleable, but As, Sb and Bi are brittle, not malleable. No sulphides
are malleable.

HARDNESS

Mineralogical “hardness” is resistance to abrasion. A harder material will scratch a


softer material, but will not be scratched by the softer one.

The empirical hardness scale devised more than a century ago by Mohs may seem
rough and ready, but it is fast and easy to use in the field or in the lab, and his benchmark
minerals were so well chosen that later quantitative measures using expensive equipment
show that hardness values 2 - 9 form a fairly uniform logarithmic progression. The hardness
standards are:

1. Talc
2. Gypsum
3. Calcite (cleavage face; rare basal faces are 21/2!)
4. Fluorite
5. Fluorapatite
6. Orthoclase
7. Quartz
8. Topaz
9. Corundum
10. Diamond

Fingernails are about 21/2.


Australian cupro-nickel (“silver”) and aluminium-bronze (“gold”) coinage is 31/2,
Glass is 51/2.
Penknife blade usually 51/2, but different grades of steel can be anywhere between 41/2 and 7!

When testing hardness, remember:

1. Try to avoid destroying good-quality crystal faces...

2. ...but use clean, flat surfaces wherever possible.

3. Wipe away powder to make sure which material is actually the one that got
scratched.

4. It can be a good idea to scratch in two directions: some materials are


harder in some directions than others (kyanite scratches at about 41/2 along the
length of blades, 7 across the blade).

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 11


5. It can also be an idea to check different crystal faces (cf. comment on calcite).

6. Porous or fine-grained materials can appear softer than they really are.

7. Intergrowths with hard minerals can appear harder than they should do.

8. The scale is not precise enough for finer divisions than half a unit to be meaningful.

Hardness can be a general indicator of composition


Most silicates are hard (> 5), as are oxides of lighter elements such as Mg, Al, Ti and
Fe. Carbonate, sulphate and phosphate minerals have hardness ≤ 5. Halides are very soft (<
4). So are most sulphides, except for those of Fe, Co, Ni and the platinum group elements.
Table C5 in Appendix C gives an overview for common species.

COLOUR

Minerals can appear coloured if different wavelengths of visible light are absorbed
or reflected to different degrees.
The colour of a mineral may be due to :
(i) The major chemical elements that are in its crystal structure.
(ii) Trace element impurities in its crystal structure
(iii) Tinting or light scattering by small inclusions or precipitates that are scattered through it.

In translucent or transparent materials without grain boundaries, you can see quite
deep below the surface, and effects (ii) and (iii) can be quite strong. In fine-grained materials
or powders, this is not the case, and only the absorption colour from (i) is discernible.
Therefore, the colour of the fine powder, or streak, is fairly constant even when a mineral
shows a wide range of tints due to trace impurities or inclusions. The streak will only show
the colours from major or significant minor elements. For instance, quartz or fluorite can be
any colour of the rainbow in hand specimen, but the streak is always white! Slightly
translucent “opaque minerals” may give a coloured streak from a black hand specimen (eg
red streak from black hematite). Some opaque metallics with a strong reflectance colour may
look different as a streak: gold, pyrite and chalcopyrite are all yellow metallic, but while gold
gives a yellow streak, pyrite and chalcopyrite do not: their streaks are noticeably different
shades of greyish or greenish black!

You should always look at both the hand specimen colour and the streak colour.

A streak can be obtained by grinding a specimen against a harder surface. Streak


plates have a Mohs hardness of about 6, so harder mineral will show you the streak colour of
the plate, not the mineral! Scratch them on something very hard (non-precious corundum
crystals make excellent streak plates for hard species).

If absorption is strong enough, even a “transparent, coloured” mineral can look


opaque and black in hand specimen. Looking at a small thickness in a grain mount, streak
powder or a thin section can often show a much more informative colour.

The colour can give you an idea of what chemical elements are in the mineral:

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 12


Blackish: check a grain mount, thin section or streak!
Black/Dark Blue/Dark Brown: Fe2+ and Fe3+ together
Blue (bright): Cu2+, V4+
Green (pure or bluish): Cu2+, V3+, Cr3+, Ni2+
Green (olive, murky): Fe2+
Yellow/Orange (bright): Cr6+, U6+, V5+
Yellow/Brown (murky): Fe3+
Red/Brown: Fe3+
Red/Purple: Mn3+, Co2+
Pink: Mn2+, Co2+

Not all elements produce a visible colour. The geochemically common elements Na,
Mg, Al, Si, P, S, Cl, K, Ca and Ti are all colourless inthemselves. Sometimes, combinations
of elements have an effect on colour, though. For instance, Ti enhances the browns due to
Fe3+ in mica and amphibole. Also, Fe3+ is redder bonded to O2-, but yellower bonded to OH-.

Colour can appear because of electronic effects that do not relate purely to one single
element. While silicate minerals are electrical insulators, many oxides and sulphides are
semiconductors, in which visible light of some wavelengths can be absorbed to move
electrons around inside the crystal. The less energy required (smaller “electronic band gap”),
the more of the visible spectrum is absorbed:

Decreasing band gap Mineral Colour when pure


↓ Sphalerite, Rutile off-white
↓ Orpiment yellow
↓ Realgar, Cinnabar transparent red
↓ Hematite nearly opaque (red when thin)
↓ Stibnite opaque (slightly transparent to
↓ very deepest red and IR!)
↓ Galena totally opaque in visible,
↓ transparent in IR
↓ Silver metal: totally opaque in visible
and in IR.

The following are examples of colours which are not from major elements in the
crystal, and do not show in a streak:

“Blue” of blue quartz (eg, from Broken Hill):


Preferential dispersion of red light (Rayleigh scattering) by submicroscopic
inclusions in colourless quartz

Iridescence of precious opal:


Diffraction by regularly spaced micron-scale spheres of cristobalite.

Intense blue and violet in halite:


“Colour centres”: a tiny fraction of chloride anions are replaced by electrons(!)

Pink in halite:
Halophilic (“salt-loving”) bacteria and algae included inside the crystals.
Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 13
Green, violet etc in fluorite:
Colour centres and trace impurites, mainly rare earth elements.

LUSTRE

The “lustre” of a mineral is a description of the quality of reflected light. The


following terms relate to the mineral’s refractive index (RI). The RI is the ratio of the speed
of a light ray in the material, to the spped of light in vacuum. A high RI means that the light
propagates more slowly. The bigger change of RI on going between the crystal and a low-RI
medium such as air also means that more light is reflected at the surface, and that light
entering the crystal at an angle is refracted more, and is more likely to be reflected around
inside the crystal. Hence, materials with very high RI (diamond, n = 2.4) looks more
“sparkly” than low-RI materials such as glass or quartz (n = 1.5). Very high RI (n > 2.5)
often correlates with strong absorption as well as strong reflection: such materials are
metallic in appearance.

Vitreous (= Glassy): quality of reflection similar to glass, like most


transparent minerals. Transparent, low or moderate
refractive index. Subvitreous is the insubstantial look
of minerals with very low RI.

Adamantine (= Diamond-like): much more intense “sparkle” due to much higher


refractive index (n more than about 1.8). Some
adamantine minerals can look as though they have
been coated with a thin metallised film!
Subadamantine is between adamantine and vitreous.

Submetallic: opaque, high refractive index and reflectance,


but not as reflective as polished metal.

Metallic: opaque and very highly reflective.

The following lustre terms relate to the texture of the mineral and directionality of reflection:

Pearly: higher than normal reflectivity and maybe irridescent


interference colours from stacks of parallel cleavage
fractures in transparent minerals with perfect cleavage.

Silky: from transparent minerals in parallel fibres

Greasy: often implies a very thin film of surface alteration

The following are influenced by colour:

Pitchy: vitreous or adamantine and black in hand specimen

Resinous: vitreous or adamantine and yellow/orange/brown

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 14


And any mineral can also appear:

Dull: diffuse light scattering from fine-grained rough


surface.

Adamantine lustre indicates high refractive index (RI) and therefore strong interaction
between the atoms of the crystal and a light ray travelling through it. This may be because of
very strong bonding (diamond, rutile) or the presence of heavy elements with lots of
electrons per atom (Pb in cerussite, Ag in chlorargyrite).

(Sub-)Metallic lustre indicates narrow (or non-existent) electron band gap, ie the material is
a semiconductor or a metal, and probably belongs to the native element (eg copper), sulphide
(eg pyrite) or oxide (eg magnetite) groups.

THIN SECTION: PLANE-POLARISED LIGHT

The standard thickness of a petrographic section is 30 microns (0.03 mm), but thicker
sections are often used (i) to identify low-birefringence minerals (see below), (ii) study fluid
inclusions inside mineral grains. You can also make temporary grain mounts of non-standard
thickness very quickly by putting loose crushed grains of a mineral on a slide in a drop of oil
or water.

The most unique feature of a petrological microscope is that the light source passes
through a polariser. Instead of vibrating in random directions transverse to the direction of
travel, the electric field component of the light is restricted to oscillate in one plane only.

For crystalline materials with less than cubic symmetry (ie, for most minerals), the
refractive index and absorption behaviour vary depending on the relative orientation of the
electric field of the light and of the crystal structure. Rotating a crystal in a beam of plane-
polarised light therefore provides much more information than looking at it, motionless, in
unpolarised light. Combined with observation between crossed polarisers, enough
information can often be obtained not only to unambiguously identify the family to which the
mineral belongs, but to pin down which species it is.

When looking at a thin section, you should always remember that a crystal can show
distinctive features in all three dimensions, but that each grain in a thin section is only a two
dimensional slice through those three dimensions. Looking at one grain rarely tells you the
full story. You need to look at several grains of the same mineral in order to reconstruct what
three-dimensional behaviour corresponds to the various two-dimensional slices.

The main optical features of a mineral that can be identified in plane polarised light
are relief, colour and microstructure.

Relief:
The edges and internal fractures of a grain stand out much more clearly (“high
relief”) if its refractive index is very different from the adjacent material, which might be
another mineral, epoxy (n = 1.54), oil, water (n = 1.33) or air (n = 1.00). Conversely, low
relief grains can appear very ghostly, vanishing into their surroundings. Immersion of a

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 15


grain in oils of different RI is still used to measure refractive index accurately (±0.0001). An
exact match causes all sharp edges and boundaries around the grain to vanish! A little
familiarity with different known minerals, will allow you to estimate RI to within ±0.1 or so
with just a quick look down the microscope.

Relief may be positive (RI higher than surroundings) or negative (RI lower than
surroundings). This may be checked by the Becke Line test (Fig. 6). Find a grain whose
margin wedges out with no bumps or chinks. Focus as sharply as you can, raise the condenser
(slightly convergent light) and partially close the aperture (less glare and bigger depth of
focus). Now, increase the distance between specimen and objective lens, watching the edge.
As it goes out of focus, the blurred image of the edge will split so that either the inside is
bright and the outside is dark, or the other way round. If the bright side moves inwards as
the objective moves away, then the relief is positive. Otherwise, it is negative.

RI varies with light vibration direction through no-cubic crystals (birefringence). If


this variation is very large, you can detect it in plane polarised light, because the relief of the
crystal will change from a higher extreme to a lower extreme as the stage is rotated. This
effect is known as “twinkling”, and is shown by calcite. Smaller birefringence values can be
measured accurately using crossed polarisers (see below).

Colour:
Because of the thinness of the section, many minerals that show a colour in hand
specimen will be colourless in thin section or nearly so. Conversely, many “black” hand
specimens show distinctive colours in a section. For instance, augite and aegirine are both
“black” clinopyroxenes in hand specimen, but augite is usually nearly colourless with a faint
brown tint in thin section, whereas aegirine shows strong green and yellow colours.

Like RI, light absorption can change depending on the vibration direction of the light
relative to the crystal structure. As an edge-on flake of biotite mica is rotated, the light is
vibrating parallel to the strongly-bonded silicate sheets of the structure in one orientation, but
across the weakly-bonded interlayers in the orientation at 90o. When light is vibrating parallel
to the layers, the RI is higher and the absorption colour is a deeper brown than when the
crystal is turned through 90 o. The wavelength of absorption may also change with absorption
direction: a hornblende may be green-blue if light is vibrating in one direction through the
crystal structure, but yellow when the crystal is turned through a right angle.

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 16


This variation in colour with vibration direction is called pleochroism, and is very
useful for mineral identification (Fig. 7). It is occasionally visible in hand specimens since
ambient slight is usually slightly polarised: large, fresh cordierite grains may change from
deep blue or violet to pal eyellowish-grey as they are rotated while held up to the light.
However, materials that are transparent enough to show hand-specimen pleochroism will be
colourless in thin section!

The light-polarising materials polaroid is actually a strongly pleochroic plastic,


whose colours are nearly colourless for one vibration direction and nearly opaque black for
the perpendicular vibration direction.

Microstructures:
Many minerals show distinctive internal inhomogeneities which can help with
identification. Examples include:

Habit:
Can you reconstruct a three-dimensional crystal shape from the two-dimensional
sections? Is the crystal equant, fibrous, prismatic, or tabular/platy? If equant, is it a cube,
octahedron or some other shape? Does it have perfect flat crystal faces (euhedral), some
rough faces (subhedral), or is it a shapeless blob (anhedral)?

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 17


Cleavage traces:
Parallel fractures, opened up during section polishing, which indicate a plane or
planes of particularly weak bonding in the crystal structure. Cleavages are useful both for
identifying minerals (eg two cleavages at right angles = pyroxene, two at 60o = amphibole)
and for working out which directions in a grain are the crystallographic x, y and z directions.

Note that cleavages only open up if they are nearly vertical in the thin section. Cleavage
planes that dip shallowly (< 40o or so) are very unlikely to be visible.

Fractures:
Curved, conchoidal fractures, as seen in olivine, are also useful since they indicate
that there are no particularly good cleavages.

Twin planes:
Mistakes during growth or later deformation can cause parts of a crystal to be oriented
differently relative to the rest of it, in a crystallographic controlled fashion. The two or more
twin individuals are coherent, so the structures are bonded together at the atomic level,
without a normal grain boundary. Nevertheless, the apparent surface texture, reflectivity,
relief, pleochroic colour and birefringence of the different twin domains can all be different.
The orientation and style are all important in identification: simple twins in orthoclase,
repeated lamellar twins in plagioclase, cross-hatched twins in microcline and anorthoclase.
The lamellar deformation twins of calcite and dolomite are on different planes, and are
arranged differently relative to traces of the cleavages. Calcite twins are parallel to the edges
and long diagonal of the rhomb, while dolomite twins are parallel to the long and short
diagonals but not the edges.

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 18


Presence of twinning may eliminate a whole mineral group: simple twins occur
frequently in monoclinic pyroxenes and amphiboles but not in orthorhombic ones!

Twins can occasionally be seen in hand specimen or thin section under plane
polarised light , but are most obvious in crossed polars (below).

Zoning:
Change in composition during growth can cause distinctive colour patterning in plane
polars, but is more obvious in crossed polars (below). Patterns may include sector zoning
(different faces show different colour/birefringence), concentric zoning (core and rim are
different), and oscillatory zoning (colour changes back and forth several times during
growth).

Exsolution lamellae:
Some minerals are confined to narrower composition ranges as they cool down, and
components that have become unstable precipitate as thin lamellae of a second mineral in the
first (eg Ca-rich pyroxene in Ca-poor pyroxene matrix).

Inclusions:
Some minerals are always quite clean, and devoid of internal precipitates. Some (eg
quartz, halite) often contain small inclusions of fluid, which may have both liquid and vapour
phases visible. Some (eg K-feldspar, cordierite) often contain a dusting of clay mineral
alteration products.

Intergrowth:
Some minerals readily grow intergrown through their neighbours. A coarse-scale
example is the Swiss-cheese like “sieve texture” of staurolite with fingers of quartz growing
though it. Fine-grained, wormy examples where two or more minerals have intergrown
include quartz-feldspar “myrmekite” and symplectites from mineral breakdown in dry,
high-grade metamorphic rocks (eg orthopyroxene-spinel-plagioclase symplectite by
breakdown of a garnet).

Pleochroic Haloes:
Accumulated radiation damage from small inclusions of mildly radioactive minerals
(often zircon or monazite) cause a ring of tell-tale pleochroic discolouration in some
minerals: dark brown in biotite and hornblende, yellow in cordierite.

Reaction Rims:
Some minerals have telltale alteration products as rims, veins or partial replacement,
of the bulk crystal. Olivine is often partially replaced by fine-grained, low-birefringence
serpentine. Brown biotite is often interlayed with pale green chlorite to give a stripy “choc-
mint” effect. Volcanic (not plutonic or metamorphic) hornblendes are oxidised while flying
through the air red-hot, and develop a black opaque crust of iron oxides.

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 19


THIN SECTION: CROSS-POLARISED LIGHT

The insertable/removable upper polariser completely changes the appearance of the


thin section. Some grains become black. Others develop bright colours that wink in and out
every 90o as the stage is rotated. These colours can be used to estimate birefringence and
crystal orientation accurately. Additional data obtainable that are useful in mineral,
identification include sign of birefringence (length-fast/length-slow), extinction angle, and
anomalous extinction due to dispersion. More sophisticated measurements such as optic
axial angle, detailed characterisation of dispersion and reflectivity at different wavelengths
are beyond the scope of this course.

The pretty colours originate when the RI varies depending on the orientation of the
light vibration in the crystal structure (ie, the crystal is birefringent). The types of variation
that are possible depends on the crystal system as follows:

Cubic/Amorphous:
RI the same for any vibration direction. No birefringence (optically isotropic). Black
when viewed in croosed polars, for any rotation of the stage.

Tetragonal/Hexagonal/Trigonal:
RI has one extreme value for vibration || z (the “extraordinary” or ε (epsilon)
direction), but has a different extreme value for vibration anywhere in the xy
plane:“ordinary” or ω (omega) directions. RI varies smoothly from one extreme to the other
for intermediate vibration directions. There is one viewing direction (down z) where the
crystal looks isotropic because the x and y
RI values are the same. This direction is the optic axis; the crystal is uniaxial because there
is one one such direction. The “optic sign” is the sign of (ε − ω) RI’s. A mineral with most of
its strong bonds in the xy plane will usually be optically negative (eg calcite), whereas a
crystal where the bonds tend to form chains running up z will usually be optically positive
(eg quartz).

Orthorhombic:
Imagine a three-dimensional polar graph in which RI is plotted as a radius against
vibration direction. Such a diagram often appears in textbooks, and is called the optical
indicatrix (Fig. 8). The indicatrix is a sphere for a cubic crystal (RI the same in all
directions), and for a uniaxial crystal it is a uniaxial ellipsoid (a shape with a circular equator
in the xy plane but with any other cross-ection an ellipse). This is a prolate ellipsoid
(football) for optically positive uniaxials and an oblate ellipsoid (M’n’M) for the optically
negative case.

If the symmetry is any lower, the indicatrix is a triaxial ellipsoid. There is a


maximum RI called γ (gamma), for a vibration direction somewhere at right-angles to it is a
minimum RI α (alpha), and for the vibration direction perpendicular to both of these is the
third principal refractive index β (beta), which takes a value somewhere intermediate (not
necessarily halfway) between α and γ. If α ≈ β << γ, the crystal is nearly uniaxial positive
(optic axis ε = γ). If α <<≈ β ≈ γ, the crystal; is nearly uniaxial negative (optic axis ε = α).
For intermediate cases, there are two optic axes somewhere in the α γ plane (symmetrically
either side of α or γ). The crystal is biaxial.

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 20


For an orthorhombic crystal, the optic principal directions α, β and γ must be
parallel to the crystallographic axes x, y and z, but not necessarily in that order!

Monoclinic:
Biaxial. One of the principal axes must be parallel to the mnoclinic twofold axis y. The other
two must be in the xz plane but do not have to align with x or z.

Triclinic:
Biaxial. No exact alignment required at all between the optic indicatrix and the
crystallographic unit cell!

Description of the optical properties of a crystal is likely to include the values of the
one, two or three principal RI’s and their orientation relative to the crystallographic axes.
Each principal RI may also have a distinctive pleochroic colour associated with it.

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 21


Slow and Fast Directions in a Thin Section
Three-dimensional optical information cannot be obtained from a cross-section of one
grain: observations from several differently oriented grains of the same mineral. Light
travelling down a single viewing direction will experience only the RI’s corresponding to the
cross-section of the optical indicatrix that is normal to the viewing direction. These RI’s will
vary smoothly between a maximum value (γ‘) and a minimum value (α‘). The γ‘ RI will be
for the vibration direction in the plane of the section that is closest to γ or furthest from α..
The pleochroic colour for γ‘ will likewise be most similar to that for γ or least similar to that
for α.. Beacuse high RI means that light is travelling slowly and low RI means that light is
travelling fast, γ’ is often called the slow direction and α’ the fast direction. These are the
directions in which the electric field of the light is vibrating, not the direction in which the
light is travelling!

Isotropic Samples and Crossed Polars


Isotropic materials (cubic system crystals and amorphous materials such as most
resins, liquids and glasses) do not affect the plane of vibration of the polarised light that
passes through them. The second polariser in the microscope, above the objective lens, is
oriented so as to block light that is polarised this way, but pass light that is polarised
perpendicular to it. Therefore, isotropic materials look black in crossed polars for any rotation
of the stage, even if they are transparent in plane polarised light! Other materials only do this
if you are viewing them down an optic axis. For instance, hexagonal-shaped cross-sections of
quartz or apatite, viewed down z, will look isotropic.

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 22


Extinction and Extinction Angle
If the crystal is rotated so that one of eitherγ‘ or α‘ is parallel to the plane of light
polarisation, then light passes straight through the crystal without any change in polarisation,
as it does for an isotropic material or for a crystal viewed down its optic axis. When the
upper polariser is in position, it completely blocks light that is vibrating in its original plane,
so the crystal looks dark. Extinction of a birefringent crystal is used to locate the principal
directions of a crystal section. Using the protractor scale on the stage to measure the
extinction angle between principal optical directions and crystallographic directions such as
cleavage traces or crystal edges can distinguish orthorhombic crystals from low symmetries,
and to identify species of mineral families such as amphiboles and pyroxenes. The extinction
directions are always parallel to projections of the crystallographic axes (“straight
extinction”) for orthorhombic crystals or monoclinic crystals viewed down a direction that is
in the xz plane. As the viewing direction for a monoclinic crystal gets closer to the y
direction, the extinction positions deviate away from any indicators of crystallographic
directions in the xz plane (“inclined extinction”). The extinction angle is biggest for crystals
viewed exactly down y. Several grains of a mineral in a section must be measured before you
can have a reliable estimate of the maximum extinction angle! The way in which a single
grain can change in appearance depending on viewing direction is shown well by hornblende
(Fig. 9).

Interference Colours
When a crystal is rotated roughly halfway between extinction positions, the
interference colour due to the birefringence is at its strongest. This colour arises because
the plane-polarising light that enters the crystal travels through it not as one beam of light, but
is split into two (double refraction). In the crystal, part of the light vibrates parallel to α‘,
and the rest parallel to γ‘. The difference in refractive indices means that these two rays travel
at different speeds through the crystal: the velocity of light in a crystal is c/n, where c is the
velocity of light in a vacuum (about 3 × 108 m/s). The frequency of a given colour of light
stays the same, but the wavelengths are contracted by a factor n = α‘ or γ‘. If the wavelength
in air (or vacuum) is λ, then the wavelengths in the crystal are λ/α‘ and λ/γ‘, and after
passing through a thickness of crystal t, there is a phase difference between the waves of
2πtγ‘/λ - 2πtα‘/λ = 2π/λ × t(γ‘-α‘) = 2π/λ × tδ, where δ is the birefringence and tδ is called
the path difference or retardation (constant for all λ). Because of the phase lag, the
resulting wave that emerges from the top of the crystal is no longer plane-polarised but
elliptically polarised.

It can be shown (Appendix B) using standard trigonometrical formulae that when this
elliptically polarised light is filtered through the crossed polariser, the final intensity varies as
sin22θ sin2(ψ/2), where θ is the angle between the first polariser direction and a principal axis
of the crystal and ψ is the phase shift on exiting the crystal, ψ = 2π/λ × tδ. The sin2θ term
describes the

fading to extinction every 90o as the stage is rotated. The sin(ψ/2) term means that as
retardation increases, a colour gradually becomes more intense and then fades again
periodically (Fig. 10).

Higher-frequency (shorter-wavelength, bluer) colours vary faster than low-frequency


(longer-wavelength, redder) ones. For zero pathdifference, no colours have any intensity and
the specimen looks black. As path difference increases, all wavelengths increase in intensity,
and the colour is no longer black but is an ever-lightening grey. With even larger retardation,

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 23


blue starts to decrease before the rest of the spectrum, giving yellow and orange transmitted
colours. As green light gets filtered out as well, the residual interference colour becomes red.
Eventually, a very intense magenta-purple colour occurs, called the sensitive tint. This is a
distinctive, particularly lurid hue, because the colour that is filtered out is the shade of green-
yellow to which the human eye is most sensitive, while red and blue

are allowed through the second polariser quite strongly. The sensitive tint marks the
end of the first order of interence colours, at a retardation of about 550 nm.
Increasing retardation produces a second order of bright rainbow colours,
finishing in a purple similar to the sensitive tint but less intense, then a slightly
washed-out rainbow of third order colours, and increasingly diffuse pearly green-
pink-grey colours from higher orders. Whereas the first and second order colours in
Fig. 10 are shown reasonably accurately, the third order is not, since the diagram uses
only three distinct colours whereas the real interference colour has contributions from
every possible wavelength. The sequence of interference colours is universal, and is
shown (usually with poor colour reproduction) on the Michel-Lévy Chart. The
sequence of interference colours is universal, and may be seen in nature in a variety of
situations where light gets reflected and partially polarised by multiple parallel
surfaces.

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 24


Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 25
The colours in mother-of-pearl, in a laminated car windscreen or on a pool of oily water are
examples.

Note that colour is a function of thickness × birefringence, so:

(i) If you know the thickness (eg 30 μm standard section), you can calculate the birefringence
rather accurately (or look it up on the Michel-Lévy chart).

(ii) If you know the birefringence, you can calculate the thickness equally accurately!

Some cautionary points:

1. Is your background (the glass slide, usually) black in crossed polars? If not, then the
orientation of your lower polariser is probably set wrongly, and your colours will be
incorrect.

2. You should note that interference colours are distorted if the mineral has a noticeable
absorption colour. This may be blatant (biotite brown in crossed polars) or subtle (unusually
bright greens and yellows for epidote).

3. The birefringences quoted on the Michel-Lévy chart are the maximum values, applicable
to the maximum-birefringence viewing direction for a mineral (down ω for a uniaxial
mineral, down β for a biaxial). Other directions will show lower-order colours.

4. Is your section thickness 30μm ? Do known minerals show their usual range of colours?

5. Is your unknown grain the full thickness of the slide? Very small grains, or randomly
oriented grains stacked on top of one another, will show low-order interference colours.

The usefulness of birefringence and relief to identify different minerals and mineral
groups is shown in Fig. 11. This shows most of the minerals that it would be useful for you to
know, and should help kick-start mineral identification in the microscope.

Appendix C provides detailed supporting data for this figure. The symbols used in the
figure are tabulated in Table C3. Table C5 summarises physical and optical properties for
most of the minerals in Figure 11 as well as some additional ore minerals, which should make
hand-specimen or thin-section identification even easier for the commonest minerals and
mineral groups.

Anomalous Interference Colours: Dispersion


Some minerals with inclined extinction do not give a clean black at a well-defined
extinction position, but “blink” between a very dark blue (“Berlin blue”) and a dark brown
colour, neither of which is part of the normal interference colour scale (see below). This is
because the extinction angle changes with wavelength (dispersion of extinction angle), and
is different at the red and blue ends of the spectrum. Examples are Ca-rich silicates such as
calcic plagioclase and clinozoisite, and Fe-rich silicates such as chlorite.
Minerals with straight extinction can also show these dark blue and brown colours if
they have low birefringence (1st order grey) which varies significantly with wavelength
(dispersion of birefringence) . A “paler grey” interference colour for blue light than for red
Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 26
(ie higher birefringence for blue light) will result in a blue tint and the opposite situation in a
brown tint. These colours do not change on rotating the stage, except to go into extinction.
Examples are zoisite and melilite (both Ca-rich silicates).

Adding and Subtracting Retardation: The Gypsum Plate


What happens if we look through sections of two birefringent materials on top of one
another? For simplicity, suppose that the extinction directions of the two are aligned parallel.
However, α’ of the first may be parallel to α’or to γ’ of the second one. If both α’ directions
are parallel, then the phase shift in the second slice is in the same direction as the phase shift
from the first slice, and the total retardation is the sum of the retardation for both slices. On
the other hand, if α’ of the first slice lines up with γ’ of the second slice, then the wave that
was phase-lagged vibrating along the slow direction of the first slice is able to catch up by
vibrating along the fast direction of the second. The total retardation is the difference of the
retardation of the two specimens. It is as if retardation has a positive or negative sign. If the
retardation of the first sample is called “positive”, then the sign of the second is positive if its
fast and slow directions point the same way, but negative if they are the wrong way round.

A “second specimen” can be inserted into the microscope in the form of the gypsum
plate (or sensitive tint plate), which is a cleavage slice of gypsum exactly thick enough to
give a sensitive tint interference colour. This is fitted into a holder, with the slow direction γ’
of the gypsum indicated. Looking at a specimen through this will change the interference
colour by exactly one order, upwards or downwards depending on whether the two
retardations are adding or subtracting. The colour depends on the absolute magnitude of the
retardation, so “-1st” order colours are the same as first order ones.

There are three major uses of the gypsum plate:

1. Checking the order of an ambiguous interference colour


For instance, a yellow colour may be first, second or third order. The gypsum plate
allows these to be distinguished in a few seconds. Look at the colour without the plate, with
the specimen at 45o to extinction so that it is as bright as possible. Then insert the plate. The
colour will go up or down by one order. Rotate sample 90o. The order of the colour will
change in the opposite direction.

Real Order Subtraction Addition


1st 1st 2nd
2nd 1st 3rd
3rd 2nd 4th

Since only 2nd and 3rd order colours are bright rainbow colours, whereas 1st order are dingy
grey-yellow-orange-brick red and 4th order are washed-out green and pink, it is easy to tell
these apart.

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 27


Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 28
2. Using the fast and slow directions to identify a crystal
For some minerals with a prismatic habit, an important identification feature is
whether the slow direction runs along the length or across the length of the crystal. This is
very fast to check by lining up the length and width of the crystal with γ’ of the gypsum plate
and seeing whether the retardations add or subtract. Sillimanite and andalusite are two
different polymorphs of Al2SiO5 that can both form prismatic crystals, but sillimanite is
length-slow (Fig. 12), but andalusite is length-fast. Cleavage can also help with these
particular minerals since sillimanite has very good cleavage parallel to only one diagonal of
the “square” prism, while andalusite has less distinct cleavages parallel to both diagonals and
one pair of edges. Some length fast/slow information has been included in Appendix C to
keep the tables compact. Note that this can be dependent on both composition (glaucophane
is different from its iron-rich relatives) and habit (a mineral with ‘length-slow’ prisms will
have ‘length-fast’ plates!). You may need to look up minerals in a reference like “DHZ” (see
below) to see exactly where α, β and γ directions are relative to cleavage and crystal forms.

3. Using the fast and slow directions to orient a known crystal


Quartz is optically positive, so in any section of a quartz grain, the vibration direction
with largest RI (the slow direction) will be the direction closest to the z axis. Inserting the
gypsum plate will convert the 1st order grey colours of quartz to either 1st order red
(subtraction colour) or 2nd order blue (addition) depending on whether the z axis is pointing
perpendicular or parallel to the γ’ direction of the plate.

TIPS ON REFERENCE BOOKS AND WWW SITES

The main text of this handbook concludes with some general and some Canberra-
specific hints on where to get more information. There are many books which can supply you
with more detailed data on minerals. Most reference books describe more species than you
are likely ever to see. The academic texts are likely to include minerals that are rock-forming
in rare rock types, and give very detailed technical data that is beyond the scope of this
course, but are good on microscopic properties and mineral associations. The popular guides
are more likely to show minerals that are visually pretty but very rare, including ore minerals,
and include hand-specimen properties but little or none of the information that you need for
microscopic work. Guides are also likely to be bad translations from German or Czech
originals, so some terminology and names may not be quite right. Nevertheless, both types of
book have their place. Note that guides are much more useful if they group minerals by
streak rather than hand specimen colour! We’ll lend a few of our personal books to the lab
for the course, but if you want your own, good places to look for cheap 2nd hand books
include Lifeline Book Fairs (EPIC, Mitchell, twice yearly), and shops such as Gilbert’s
(Civic), Canty’s (Fyshwick) and Booklore (Lyneham).

A book which is pretty much the ‘Bible’ for properties of rock-forming minerals is
W.A. Deer, R.A. Howie and J. Zussmann’s Rock-forming Minerals (‘DHZ’ for short), either
in its multi-volume form or in the single-volume student edition (Introduction to the...). W.S.
McKenzie’s Atlas of rock-forming minerals in thin section (with C. Guilford) and A color
atlas of rocks and minerals in thin section (with A.E. Adams) are also fantastic.

There are a variety of excellent WWW sites out there which provide mineral data,
often via searchable databases. Particularly recommended are the following:

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 29


http://www.mindat.org Detailed localities, discussion forums
http://www.webmineral.com Mineral properties
http://www.minsocam.org Mineralogical Society of America site with excellent
articles, pdf’s of ‘Handbook of Mineralogy’ silicate
pages.
http://un2sg4.unige.ch/athena/mineral/search.html
Database searchable using combinations of elements.
http://database.iem.ac.ru/mincryst/search.php
Russian crystal structure database.
http://www.webelements.com WWW Periodic Table with loads of additional element
data.

WWW ‘atlases’ with micrographs in thin section are easy to find by searching. For examples,
try:

www.geolab.unc.edu/Petunia/IgMetAtlas/minerals/minerals.html

ww.uwgb.edu/dutchs/petrolgy/thinsect.htm

http://sorrel.humboldt.edu/~jdl1/petrography.page.html#anchor748621

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 30


Appendix A: BASIC CRYSTALLOGRAPHY
Crystal Symmetry
Most mineral species are crystalline materials. This means that their atoms are
arranged periodically in three dimensions. A crystal does not necessarily have to have flat
faces, although these are a common feature of crystals. The crystal structure of a mineral
can be described as a pattern of atoms (“motif”) in a building block of a particular shape and
size (“unit cell”). The unit cells then stack like a pile of bricks through the crystal, and repeat
regularly (translational symmetry). Shortest distances between atoms are typically 1 - 2 Å
(Ångström units) = 0.1 - 0.2 nm = 1 - 2 × 10-10 m. The edges of a unit cell are typically a few
Å. A 1mm cube of halite (table salt) has about 2 million unit cell repeats along each edge.

Translationally symmetry means that there is a set of vectors called lattice vectors
that relate any chose point in the crystal with other points that are translationally equivalent
(same environment, in the same orientation). Any three independent lattice vectors can be
taken to define the edges of a unit cell (which is always a paralleliped). These vectors, the
crystal axes, are usually called a, b and c. The lengths of the edges (a, b, c) and angles
between them (α = b^c, β = c^a, γ = a^b) are called unit cell parameters. Note that the three
vectors define a coordinate system that is like the familiar Cartesian coordinates, except that
in general, a, b and c could all be different, and none of the angles need to be right angles.
The mathematics involved in calculated angles and distances can therefore be much more
complicated. Whatever point in the structure is taken as a starting position, displacement by
any combination of integral numbers of a, b and c will takeyou to a point that is
translationally equivalent. The set of all such points forms a three-dimensional grid called the
crystal lattice (Fig. A1). It is important to note that lattice points relate only to each other,
and are not tied to any particular arrangement of atoms in the crystal structure!

There are a restricted range of combinations of rotation axis and reflection plane
symmetry elements that are compatible with the translational symmetry of a crystal (Fig.
A2).

Symmetry elements may force some lattice vector lengths to be equal, and/or some
angles between vectors to be 90o or 60/120o. It is usual to choose unit cell edges that
highlight this symmetry, since this makes geometrical calculations much easier. Sometimes,
this requires that the unit cell is bigger in volume than the smallest possible “primitive” cell.
For example, three different arrangements of lattice points are compatible with the symmetry
of a cube, but only one of these can have a primitive cubic unit cell. Choosing a cube-shaped
cell for the other two results in a doubly primitive body-centred cubic lattice (extra lattice
point at [1/2 1/2 1/2]) or the quadruply primitive face-centred cubic lattice (extra lattice
points at [0 1/2 1/2], [1/2 0 1/2] and [1/2 1/2 0]). It is possible to choose smaller-volume primitive
cells in both cases, but the geometries of such cells are much less convenient (Fig. A3).

The most symmetrical choices of crystal axes cause the possible types of crystal
symmetry to fall naturally into seven types called crystal systems (Fig. A4). The symmetry
of the crystal system plays a major role in determining the optical properties of a mineral.
Within each system, there are varieties called crystal classes (or “point groups”) which are
determined by whether mirror planes or a centre of symmetry are present, whether a high-
symmetry axis has perpendicular twofold axes, or is not a normal rotational axis but a roto-
inversion axis. The seven systems break down into 32 classes altogether. Symmetry
variations between classes often show up in crystal habit: pyrite, sphalerite and galena belong
to different cubic classes, and this can often be seen from the faces shown on their crystals.
Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 31
Different crystal systems also have different ranges of lattice types that are
compatible with them, The three lattices of the cubic system have already been mentioned. In
total, there are 14 such distinguishable Bravais lattices. These play some role in influencing
crystal habit, but in general are not important in determining macroscopic properties. They
play a much more important role in describing the crystal structure at the atomic level. The
same is true of the space group symmetry of the crystal. Macroscopic rotation axes and
mirror planes when examined at the atomic level may have fractional lattice-vector shifts
associated with them (screw axes and glide planes). These break down the 32 crystal classes
into a total of 230 more subtle complexions of symmetry. Although Bravais lattices and space
groups are beyond the scope of this guide, some familiarity with them is needed in order to
understand how actual crystal structures relate to macroscopic symmetry, some aspects of
diffraction by crystals, and some spectroscopic and vibrational properties of minerals.

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 32


Planes and Directions
Flat, shiny faces with sharp edges and corners are the most widely known feature of
crystals. These occur because of the regular translational periodicity. Whatever set of lattice
vectors are chose as unit cell edges, each flat face can be constructed from regular stepped
arrays of unit cells. For a given shape of unit cell, the angle between two such faces is always
constant even if the relative sizes of the faces vary from crystal to crystal. Therefore, the real
symmetry of a crystal is more reliably judged by angles between faces, not the apparent
macroscopic symmetry of the crystal, which may be distorted due to crowding by other
crystals or gradients of concentration, temperature and so on when the crystal was growing.

Like the unit cells in a crystal, crystal planes repeat regularly. The spacing of a set of
planes will be constant for a particular orientation of plane. Miller indices are the standard
way of labelling such planes. These are sets of three numbers (hkl) such that h, k and l are the
number of equivalent planes that cut one repeat of the a, b and c axes respectively (Fig. A5).
In general, h, k and l are always integers. Intersection on -a, -b or -c implies a negative index,
which is usually indicated compactly by an overbar (-1 = “bar-1” ≡ -1). A set of planes that is
parallel to an axis has a corresponding index of zero. Small values of (hkl) correspond to
large plane spacings, and often to important crystal faces. For describing crystal morphology,

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 33


Miller indices are usually reduced to their lowest common denominator. For instance, some
faces of a cube in the cubic system are (100), (010), (001) and (-100). When the exact plane
spacing is important, as when discussing X-ray diffraction, multiples of these such as (020),
(003) may be used.

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 34


Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 35
Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 36
The symmetry of a crystal may mean that a plane set (hkl) is related to other planes in
different orientations. The spacing will be the same, and there will be specific angular
relationships between the faces. The six faces of a cube are examples. The resulting crystal
form, the set of all related faces, can be indicated using braces (curly brackets) around the
hkl for one member of the form: {100} or {010} refers to all six cube faces simultaneously.
Similarly, the eight faces of an octahedron can be referred to collectively as {111}. Note that
the exact form referred to depends on the crystal class. In two out of five cubic classes, {111}
refers only to a tetrahedron since (111) is related only to (1-1-1),( -11-1), and ( -1-11). The four
other “octahedral” faces constitute a second tetrahedral form {-111} in these classes.
Sphalerite often shows crystals where the two sets of tetrahedral faces are different sizes and
have different surface textures and lustres.

The edges between crystal faces are examples of crystallographic directions. These
are vectors, and can be described using the unit cell edges as a coordinate system. Direction
vectors are distinguished from Miller indices by being placed in square brackets [uvw].
Pointed brackets <uvw> indicate a “star” of symmetry-related direction vectors.
It is possible to check whether a direction is a possible edge of a face (ie, whether the
direction and face are parallel) by taking the scalar product (hkl).[uvw] = hu+kv+lw. This
value is zero if the two are parallel. A set of faces which are all parallel to one directionis
called a zone.

Plane spacing and diffraction


The spacing of planes (hkl) can be calculated from the Miller indices and the unit cell
parameters, although the algebra gets more complicated as the symmetry of the unit cell
becomes lower. For the cubic system (the simplest case), the relationship is dhkl =
a/√(h2+k2+l2), where a is the edge length of the unit cube and dhkl is the spacing of the planes.
For lower-symmetry minerals, the distinctive geometry of the lattice can produce a unique
series of d-spacings which serve as a “fingerprint” for the mineral. The first experimental
evidence of the translational symmetry of crystals was obtained when it was found that X-
rays (electromagnetic radiation with Ångström wavelengths) were strongly reflected from
crystals only when the crystal was oriented so that the radiation hit planes (hkl) at an angle θ
such that the Bragg Equation nλ = 2 dhkl sinθ was satisfied for some integer n and X-ray
wavelength λ. At such an orientation, the reflections from successive planes are all in phase,
and interfere constructively to produce a strong reflection. Otherwise, they cancel each
other out. This phenomenon is known as diffraction. The angles at which a crystal diffracts
X-rays can be used to calculate the size and shape of the unit cell, and the intensities of the
reflections ultimately reveal information about the positions of the atoms.

The factor n arises because X-rays do not diffract only at angles corresponding to
(hkl) but can also diffract at the angles for (2h 2k 2l), (3h 3k 3l) and so on. This happens even
though it is not possible to find correspondingly place d layers in a crystal structure model.
The reason is that the X-rays see the crystal structure as a continuous cloud of varying
electron density. The detailed structure can be made up by superposing sinusoidal waves at
not just the basic frequency (plane spacing) but also twice the frequency (half the spacing),
three times the frequency (one-third spacing) and so on. Physicists will recognise this as an
example of Fourier synthesis. Musicians may think of it as similar to producing a tone from a
fundamental note and harmonics.

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 37


Conversely, when additional planes of crystal structure are present, small-index
planes may cease to diffract! A body-centred or face-centred cube has layers of structure at
height z+1/2 that are exactly equivalent to those at height z. These diffract out-of-phase with
each other so that the intensity of diffraction for (001) is zero. Such systematic absences in
diffraction patterns provide information about the lattice type.

X-ray diffraction is an extremely important technique in mineralogy. It can be used


for:
(i) Determining or refining a crystal structure.
(ii) Identifying a mineral from its d-spacing fingerprint. This includes distinguishing
between two
polymorphs of the same composition but different structure (e.g. calcite and
aragonite).
(iii) Measuring accurate unit cell parameters in order to obtain molar volume data or to
determine
the composition of a solid solution.
(iv) Quantifying the proportions of different minerals in a mixture
(v) Measuring strain and particle size
(vi) Measurements at high temperatures or pressures to obtain thermal expansivity and
compressibility data.

Examples of some crystal forms are shown in the figure on the next page.

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 38


Figure A6

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 39


APPENDIX B: DERIVATION OF CROSSED-POLARS INTENSITY
VARIATIONS GIVING RISE TO INTERFERENCE COLOURS
(primarily for physicists):

Intial plane-polarised wave has unit amplitude, and has electric field E = sin(2πz/λ)

Polariser angle is φ relative to E-W direction (crossed polar will always be N-S):

α = angle between α’ direction and E-W.

Incident wave is resolved into α’ component cos(α−φ)sin(2πα’z/λ) and γ’ component


-sin(α−φ)sin(2πγ’z/λ).

Put z = t: phases are now sin(2πα’t/λ) and sin(2πγ’t/λ) respectively.

If we substitute 2πα’t/λ = constant κ for the exit wave,

then α’ and γ’ components after exit from crystal can be written:

cos(α−φ)sin(2πz/λ + κ) and -sin(α−φ)sin(2πz/λ + 2π(γ’−α’)z/λ + κ)

phase of γ’ at a given z is ahead by 2π(γ’−α’)z/λ = 2πtδ/λ = ψ

total exit wave is now elliptically polarised, and can be described in terms of three components, the α’
wave, unshifted γ’ wave and a π/2-shifted γ’ wave:

α’ = cos(α−φ) sin(2πz/λ+ κ)
γ’0 = -sin(α−φ) cosψ sin(2πz/λ+ κ)
γ’π/2 = -sin(α−φ) sinψ cos(2πz/λ+ κ)

On passing through the crossed polariser, each of these have projected components parallel to the N-S
crossed polar:

α’X = sinα cos(α−φ) sin(2πz/λ+ κ)


γ’0,X = -cosα sin(α−φ) cosψ sin(2πz/λ+ κ)
γ’π/2,X = -cosα sin(α−φ) sinψ cos(2πz/λ+ κ)

with amplitudes:

Aα’X = sinα cos(α−φ)


Aγ’0,X = -cosα sin(α−φ) cosψ
Aγ’π/2,X = -cosα sin(α−φ) sinψ

The resultant intensity is therefore:

I = (Aα’X + Aγ’0,X)2 + A2γ’π/2,X

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 40


= (sinα cos(α−φ) − cosα sin(α−φ) cosψ)2 + (cosα sin(α−φ) sinψ)2

If the first polariser is set so that φ = 0, this becomes:

I = (sinα cosα − cosα sinα cosψ)2 + (cosα sinα sinψ)2

= (sinα cosα)2{(1 − cosψ)2 + sin2ψ}

1
= /4 sin22α.(2 − 2cosψ)
1
= /2 sin22α.(1 − cosψ)

= sin22α. sin2(ψ/2)

Whereas if the polarizer is at φ = π/2:

I = (sinα sinα − cosα cosα cosψ)2 + (cosα cosα sinψ)2

= (1 − cos2α + cos2α.cosψ)2 + (cos2α.sinψ)2

= {1 + 2(1 − cosψ)(cos4α - cos2α)}

= {1 − 2(1 − cosψ).cos2α.sin2α}

= 1 − 1/2 sin22α.(1 − cosψ)

= 1 − sin22α. sin2(ψ/2)

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 41


APPENDIX C: GEOCHEMICAL AND MINERALOGICAL TABLES
Table C1: Mineralogically important elements listed by symbol

Additional columns give atomic numbers (Z), atomic masses (M), siderophilicity index (S: see
notes, below), commonest coordination numbers to oxygen (CN), main naturally occurring
oxidation states (OS).

Element Z M S CN OS
Ag Silver (Argentum) 47 107.9 7 2 0, +1
Al Aluminium 13 27.0 1 4-6 +3
As Arsenic 33 74.9 6 4 -3, 0, +3, +5
Au Gold (Aurum) 79 197.0 8 4 0, +1, +3
B Boron 5 10.8 1 3-4 +3
Ba Barium 56 137.3 2 9-12 +2
Be Beryllium 4 9.0 1 4 +2
Br Bromine 35 79.9 n/a n/a -1
C Carbon 6 12.0 0 3 -4, 0, +4
Ca Calcium 20 40.1 2 6-9 +2
Ce Cerium 58 140.1 1 9-12 +3, +4
Cl Chlorine 17 35.5 n/a n/a -1
Co Cobalt 27 58.9 5 6 +2, +3
Cr Chromium 24 52.0 2 6 +3, +6
Cs Cesium 55 132.9 2 9-12 +1
Cu Copper (Cuprum) 29 63.6 6 4-6 0, +1, +2
F Fluorine 9 19.0 n/a n/a -1
Fe Iron (Ferrum) 26 55.9 4 6 0, +2, +3
Ga Gallium 31 69.7 5 4-6 +3
Ge Germanium 32 72.6 5 4-6 +4
H Hydrogen 1 1.0 2 1-2 +1
Hf Hafnium 72 178.5 1 6-8 +4
Hg Mercury (Hydrargyrum) 80 200.6 7 2-6 0, +1, +2
I Iodine 53 126.9 1 3 -1, +5
K Potassium (Kalium) 19 39.1 2 9-12 +1
La Lanthanum 57 138.9 1 9-12 +3
Li Lithium 3 6.9 2 4-6 +1
Mg Magnesium 12 24.3 2 6 +2
Mn Manganese 25 54.9 3 6-8 +2, +3, +4
Mo Molybdenum 42 95.9 5 4-6 +4, +6
N Nitrogen 7 14.0 0 3 -3, 0, +5
Nb Niobium 41 92.9 2 6 +5
Nd Neodymium 60 144.2 1 9-12 +3
Ni Nickel 28 58.7 5 6 0, +2
Na Sodium (Natrium) 11 23.0 2 6-12 +1
O Oxygen 8 16.0 n/a n/a -2
P Phosphorus 15 31.0 0 4 +5
Pb Lead (Plumbum) 82 207.2 5 4-12 0, +2, +4
Pt Platinum 78 195.1 8 6 0, +2, +4
Rb Rubidium 37 85.5 2 9-12 +1
S Sulphur 16 32.1 0 4 -2, 0, +6
Sb Antimony (Stibium) 51 121.8 6 4-6 -3, 0, +3, +5
Sc Scandium 21 45.0 1 6 +3
Se Selenium 34 79.0 6 3 -2, 0, +4
Si Silicon 14 28.1 0 4 +4
Sn Tin (Stannum) 50 118.7 4 6 +2, +4
Sr Strontium 38 87.6 2 9-12 +2
Ta Tantalum 73 181.0 2 4-8 +5
Te Tellurium 52 127.6 6 4-6 -2, 0, +4 , +6
Th Thorium 90 232.0 1 8 +4
Ti Titanium 22 47.9 2 6 +4
Tl Thallium 81 204.4 5 6 +1, +3
U Uranium 92 238.0 1 6-8 +4, +6
V Vanadium 23 50.9 3 4-6 +3, +4, +5
W Tungsten (Wolfram) 74 183.9 2 4-6 +4, +6
Y Yttrium 39 88.9 1 6-9 +3
Zn Zinc 30 65.4 5 4-6 +2
Zr Zirconium 40 91.2 1 6-8 +4

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 42


Notes:
1. The siderophilicity index is a concept unique to this handbook. It gives an empirical summary of
what types of mineral are formed by the element in the Earth’s crust, as follows:

S = 8: strong siderophile. Native element, intermetallic compounds. Oxy-compounds rare.


S = 7: sidero-/chalcophile. Primary ores are sulphides. Native element a common supergene
product. Oxycompounds rare.
S=6: strong chalcophile. Primary ores are sulphides. Native element a common supergene
product. Oxycompounds common in oxidised zone.
S=5: chalcophile. Primary ores are sulphides. Native element rare or non-existent.
Oxycompounds common in oxidised zone.
S=4: chalco-/lithophile. Primary mineralisation is oxide or sulphide, about equally.
S=3: weak lithophile. Primary mineralisation mainly oxycompounds, but sulphides
can occur in economically significant deposits.
S=2: lithophile. Primary mineralisation mainly oxycompounds. Sulphides
only known from strongly reducing environments such as meteorites.
S=1: strong lithophile. No minerals with bonds to sulphur are known or likely.
S=0: litho-/siderophile. Lithophile in the crust, but these elements can form intermetallic
compounds in meteorites and probably the Earth’s core.

this index is not relevant to elements which do not bond to oxygen because they only form anions.

2. Coordination number gives an indication of the size of the atom. Bigger cations have room for
more oxygen neighbours, unless there is a good reason why they can’t have them. Not all oxygen
coordination numbers and oxidation states are included. When there is more than one oxidation
state, the cordination number range refers to the oxidation state that is shown in bold.

3. Elements in negative oxidation states are usually in the form of “-ide” anions such as sulphide
(S2-), chloride (Cl-), selenide (Se2-), bromide (Br-), telluride (Te2-), iodide (I-). Carbon has the formal
oxidation state -4 in methane (CH4). Nitrogen has -3 in ammonia and compounds of the ammonium
cation, NH4+.

Complex oxyanions are formed by elements in high oxidation states. These are usually “-ate”
anions (eg nitrate, NO3-, and sulphate SO42- from N5+ and S6+ respectively). Some anions are formed
by elements in oxidation states other than their highest: these may be distinguished by being called
“-ites” rather than “-ates”. For instance, As3+ forms the arsenite anion AsO33-, while As5+ forms
arsenate AsO43-, and As3- forms arsenide As3-!

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 43


Table C2: Top 15 most abundant chemical elements in various environments

Concentrations are given in parts per million by mass. Notice the rapid fall-off down the columns
means that only a few elements are really common, but these are different in different parts of the
Earth! The carbonaceous chondrite values are those of cold, undifferentiated solid Solar System
material that retains much of its volatile C, H, S. Loss of these during formation of the Earth, and
loss of Fe and Ni to the core, put all of these elements much lower in the mantle list, which is
dominated by O, Mg and Si. Hence, upper mantle mineralogy is dominated by the Mg silicates
olivine and orthopyroxene. The relative incompatibility of Al, Ca, Na and K in mantle minerals
means that they become much more important in the crust, so the main silicates in the crust are
feldspars and their low-temperature equivalents. Note that seawater concentrates even further some
soluble species such as Cl and B, which are therefore important in evaporite minerals although rare
otherwise.

Data is taken from www.webelements.com, except the primitive mantle estimates, which are from
Palme, H. and O'Neill, H. (2004) Geochemical Estimates of Mantle Composition, in Treatise on
Geochemistry. Ed. R. W. Carlson, Elsevier, Oxford, 1-38pp.

Carbonaceous Primitive Mantle Earth’s Crust Sea Water


Chondrite
O 410000 O 443300 O 460000 O 857000
Fe 220000 Mg 221700 Si 270000 H 107800
Si 140000 Si 212200 Al 82000 Cl 19870
Mg 120000 Fe 63000 Fe 63000 Na 11050
S 41000 Ca 26100 Ca 50000 Mg 1326
H 24000 Al 23800 Mg 29000 S 928
C 15000 Na 2590 Na 23000 K 416
Ni 13000 Cr 2520 K 15000 Br 67
Ca 11000 Ni 1860 Ti 6600 C 28
Al 9300 Ti 1280 C 1800 Sr 8
Na 5600 Mn 1050 H 1500 B 4
Cr 3100 K 260 Mn 1100 Ca 4
Mn 2800 S 200 P 1000 F 1
N 1400 H 120 F 540 Si 1
P 1100 C 102 S 420 N 0.5

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 44


Table C3: Index of Minerals, Mineral Formulae and Mineral Symbols

This alphabetically arranged index includes all the minerals of Fig. 11 and the other tables
of this Appendix, plus some extra relatives and associates and the names of some mineral groups.
Some chemical end-members such as ferrosilite are rare as minerals, but may be encountered
labelling the corners of chemographic diagrams. The two/three-letter symbols are those
recommended by the International Mineralogical Association, are frequently used in the literature,
and are becoming increasingly standardised. These are the symbols used in Figure 11 of the main
text.

Some chemical formulae for a general group replace specific cations by ‘A’ (larger cation
such as Ca), ‘B’ (medium-sized cation such as Mg) or ‘T’ (small, tetrahedrally coordinated cation,
usually Si or Al). Anionic complexes have been enclosed in square brackets for clarity. The
oxidation state of Fe is +2 unless otherwise indicated.

The low-temperature zeolite and clay minerals are usually too fine-grained for identification
in thin section. However, zeolites may form large crystals as vesicle fills in mafic volcanic rocks. A
couple of the large number of zeolite species are listed to give examples of typical formulae.

Actinolite (Ca-Fe rich amphibole) Ca2(Fe,Mg)5[Si8O22](OH)2 Act


Aegirine (Na-Fe3+ Cpx) NaFe3+[Si2O6] Ae
Albite (Na plagioclase feldspar) Na[AlSi3O8] Ab
Almandine (Fe2+-Al garnet) Fe3Al2[SiO4]3 Alm
Amphibole (double-chain silicate group) A2-3B5[T8O22](OH)2
Analcime (Na feldspathoid) Na[AlSi2O6].H2O Anl
Andalusite (low-P aluminosilicate) Al2[SiO4]O And
Andradite (Ca-Fe3+ garnet, contact metam. rocks) Ca3Fe3+2[SiO4]3 Adr
Anhydrite (from burial/heating of gypsum) Ca[SO4] Anh
Ankerite (Fe equiv. of dolomite) CaFe[CO3]2 Ank
Anorthite (Ca plagioclase feldspar) Ca[Al2Si2O8] An
Anthophyllite (Mg orthoamphibole) Mg7Si8O22(OH)2 Ath
Apatite (group of Ca phosphates) Ca5[PO4]3(F,Cl,OH) Ap
Aragonite (high-P) Ca[CO3] Arg
Arfvedsonite (Na-Fe rich amphibole) Na3Fe2+4Fe3+[Si8O22](OH)2 Arf
Arsenopyrite (commonest As mineral) Fe[AsS] Apy
Augite (commonest igneous Cpx) (Ca,Na)(Mg,Fe,Al,Ti)[(Si,Al)2O6] Aug
Baryte (commonest Ba mineral) Ba[SO4] Brt
Bismuthinite (Bi sulphide) Bi2S3 Bmt
Beryl (commonest Be mineral) Al2[Be3Si6O18] Brl
Biotite (Mg,Fe-rich mica subgroup) K(Mg,Fe,Al)3[(Si,Al)4O10](OH)2 Bt
Borax (major B source) Na2[B4O5(OH)4].8H2O
Bornite Cu5FeS4 Bn
Brucite Mg(OH)2 Brc
Calcite (low-P carbonate) CaCO3 Cal
Cassiterite (commonest Sn mineral) SnO2 Cst
Celestite (commonest Sr mineral) Sr[SO4] Cls
Chabazite (zeolite) Ca[Al2Si4O12].6H2O Cbz
Chalcocite Cu2S Cc
Chalcopyrite CuFeS2 Ccp
Chamosite (Fe equiv. of chlorite, in ironstones) Fe5Al[AlSi3O10](OH)8 Chm
Chlorite (Mg-Al layer silicate group) Mg5Al[AlSi3O10](OH)8 Chl

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 45


Chloritoid (Fe-Al silicate in metapelites) FeAl2[SiO4]O(OH)2 Cld
2+
Chromite (Fe -Cr spinel) FeCr2O4 Chr
Cinnabar (Major Hg ore) HgS
Clinopyroxene (group) AB[T2O6] Cpx
Clinozoisite (Fe-poor equiv. of epidote) Ca2Al3[Si2O7][SiO4]O(OH) Czo
Coesite(high-P polymorph) SiO2 Coe
Cordierite (high-T Mg aluminosilicate) Mg2[Al4Si5O18] Crd
Corundum Al2O3 Crn
Cristobalite (high-T low-P polymorph) SiO2 Crs
Diamond (high-P polymorph) C
Diaspore (in bauxite/regolith) AlO(OH) Dsp
Diopside (Ca-Mg Cpx) CaMg[Si2O6] Di
Dolomite (Ca-Mg carbonate) CaMg[CO3]2 Dol
Enstatite (Mg Opx) Mg2[Si2O6] En
Epidote (Ca-Fe3+-Al disilicate) Ca2Al2Fe3+[Si2O7][SiO4]O(OH) Ep
Fayalite (Fe olivine) Fe2[SiO4] Fa
Ferrosilite (Fe Opx) Fe2[Si2O6] Fs
Fluorite CaF2 Fl
Forsterite (Mg olivine) Mg2[SiO4] Fo
Galena PbS Gn
Garnet (silicate group) A2+3B3+2[SiO4]3 Grt
Gibbsite (in bauxite/regolith) Al(OH)3 Gbs
Glaucophane (Na-Mg amphibole in blueschist) Na2Mg3Al2[Si8O22](OH)2 Gln
Goethite Fe3+O(OH) Gt
Graphite (low-P polymorph) C Gr
Grossular (Ca-Al garnet) Ca3Al2[SiO4]3 Grs
Gypsum (Ca sulphate) Ca[SO4].2H2O Gps
Halite NaCl Hl
Hedenbergite (Ca-Fe2+ Cpx) CaFe[Si2O6] Hd
Hematite Fe2O3 Hem
Hercynite (Fe2+-Al spinel) FeAl2O4 Hc
Hornblende (amphibole subgroup) (Ca,Na)2+x(Mg,Fe,Al,Ti)5[(Si,Al)8O22](OH)2 Hbl
Illite (mica/’clay’: low-P low-T var. of Mus) (K,H3O+)1-xAl2[Al1-ySi3+yO10](OH)2 Ill
Ilmenite FeTiO3 Ilm
Jadeite (high-P low-T Na-Al Cpx) NaAl[Si2O6] Jd
Kaolinite (Al-rich clay) Al2[Si2O5](OH)4 Kln
K-feldspar (group of polymorphs) K[AlSi3O8] Kfs
Kyanite (high-P aluminosilicate) Al2[SiO4]O Ky
Leucite (K feldspathoid) K[AlSi2O6] Lct
Lawsonite (high-P low-T Ca-Al silicate) CaAl2[Si2O7](OH)2.H2O Lws
Magnesite (Mg carbonate) MgCO3 Mgs
Magnetite (Fe2+-Fe3+ spinel) Fe2+Fe3+2O4 Mgt
Marcasite (unstable polymorph of pyrite) Fe[S2]
Mica (layer silicate group) AB2-3[T4O10](OH)2
Molybdenite (commonest Mo mineral) MoS2 Mlb
Monazite (rare earth phosphate) (Ce,La...)[PO4] Mnz
Muscovite (Al mica) KAl2[AlSi3O10](OH)2 Ms
Natrolite (zeolite) Na2[Al2Si3O10].2H2O Ntr
Nepheline (Na feldspathoid) (Na,K)[AlSiO4] Ne
Olivine (group) A2[SiO4] Ol
Orpiment (As sulphide) As2S3
Orthopyroxene (group) A2[T2O6] Opx

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 46


Pectolite (Na-Ca pyroxenoid) NaCa2[Si3O8(OH)] Pct
Pentlandite (major Ni ore) (Ni,Fe)9S8 Pnt
Phengite (mica: high-P/low-T muscovite variety) K(Al,Mg,Fe)2[Al1-xSi3+xO10](OH)2 Phg
Phlogopite (mica: Mg equiv. of biotite) KMg3[AlSi3O10](OH,F)2 Phl
Plagioclase feldspar (group) (Ca,Na)[Al1+xSi3-xO8] Pl
Prehnite (low-T Ca-Al layer silicate) Ca2Al[AlSi3O10](OH)2
Prh
Pumpellyite (low-T Ca-Fe-Al silicate) Ca2(Fe,Mg)(Al,Fe)2[Si2O7][SiO4](OH)2.H2O Pmp
Pyrite FeS2 Py
Pyrope (high-P Mg-Al garnet) Mg3Al2[SiO4]3 Prp
Pyrophyllite (low-T Al layer silicate) Al2[Si4O10](OH)2 Prl
Pyrrhotite Fe0.95S Po
Quartz SiO2 Qtz
Realgar (As sulphide) [As2]S2
Rhodochrosite (Mn equiv. of calcite) Mn[CO3] Rds
Riebeckite (Na-Fe2+-Fe3+ amphibole in ironstones) Na2Fe2+3Fe3+2[Si8O22](OH)2 Rbk
Rutile TiO2 Rtl
Scapolite (group of feldspathoids) (Ca,Na)4[(Si,Al)12O24](CO3,Cl) Scp
Serpentine (Mg layer silicate group) Mg3[Si2O5](OH)4 Srp
Siderite (Fe carbonate) Fe[CO3] Sd
Sillimanite (high-T aluminosilicate) Al2[SiO4]O Sil
Smectite (group of clay minerals) (Na,Ca)x(Al,Mg,Fe)2-3[(Si,Al)4O10](OH)2.nH2O Sm
Smithsonite (Zn equiv. of calcite) Zn[CO3]
Sphalerite (=’zincblende’) ZnS Sp
Spinel (Mg-Al spinel) MgAl2O4 Spl
Spinel group A2TX4
Spessartine (Mn2+-Al garnet) Mn3Al2[SiO4]3 Sps
2+
Staurolite (Fe-Al silicate in metapelites) Fe 2Al9[SiO4]4O7(OH) St
Stibnite (commonest Sb mineral) Sb2S3
Stishovite (very high-P polymorph) SiO2 Sti
Sulphur S
Sylvite KCl
Talc (Mg layer silicate) Mg3[Si4O10](OH)2 Tc
Tetrahedrite (Cu-Sb sulphide) (Cu,Ag)10(Fe,Zn,Cu)2Sb4S13 Td
Titanite (Ca-Ti silicate) CaTi[SiO4]O Ttn
Topaz (Al silicate with F) Al2[SiO4](F,OH)2 Tpz
Tremolite (Ca-Mg rich amphibole) Ca2Mg5[Si8O22](OH)2 Trm
Tridymite (low-P high-T polymorph) SiO2 Trd
Tourmaline (group of borate-silicates) Na(Fe,Mg,Li,Al)3Al6[Si6O18][BO3]3(OH,F)4 Tur
Wolframite (group of W ores) (Fe,Mn)WO4
Wollastonite (CaSiO3 pyroxenoid) Ca3[Si3O9] Wo
3+
Vesuvianite (Ca-Mg-Al disilicate) Ca19(Mg4Al8)Fe [Si2O7]4[SiO4]10(OH,F)8 Ves
Zoisite (epidote group) Ca2Al3[Si2O7][SiO4]O(OH) Zo
Zircon (commonest Zr mineral) Zr[SiO4] Zrc

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 47


Table C4: Common minerals grouped by paragenesis
IGNEOUS ROCKS mafic Olivine (Mg-rich), Orthopyroxene, Clinopyroxene
(augite), Plagioclase (calcic), Ilmenite, Magnetite
felsic Quartz, Plagioclase (sodic), K-feldspar, Biotite,
subalkaline Amphibole (hornblende), Titanite
peralkaline Nepheline, Feldspar (albite, microcline), Clinopyroxene
(aegirine), Amphibole (arfvedsonite), Titanite
METABASITES greenschist Epidote, Chlorite, Albite, Amphibole (actinolite)
facies
amphibolite Amphibole (hornblende), Plagioclase (sodic-calcic),
facies Garnet (almandine)
granulite facies Orthopyroxene, Clinopyroxene (diopside), Plagioclase
(calcic), Garnet (almandine-pyrope)
blueschist facies Chlorite, Epidote, Amphibole (glaucophane),
Clinopyroxene (jadeite), Lawsonite, Garnet (almandine-
pyrope), Aragonite
eclogite facies Clinopyroxene (omphacite), Garnet (pyrope), Kyanite,
Zoisite, Rutile
METAPELITES greenschist Quartz, Muscovite, Chlorite. Biotite at high-T end of
facies range. Chloritoid if Fe-rich.
amphibolite Quartz, Muscovite, Biotite, Garnet (almandine).
facies Aluminosilicate (kyanite at higher P & lower T,
anadlusite in low-P equivalents, sillimanite at highest T)
granulite facies Quartz, K-feldspar, Sillimanite, Garnet (almandine) +
maybe Biotite, Cordierite
CALC-SILICATES amphibolite Calcite, Amphibole (tremolite), Clinopyroxene (diopside),
facies Garnet (grossular-andradite), Epidote/Clinozoisite/Zoisite,
Vesuvianite
granulite facies Calcite, Clinopyroxene (diopside), Olivine (Mg-rich),
Scapolite, Wollastonite
Cu-Pb-Zn ORES primary chalcopyrite (Cu, Fe), sphalerite (Zn), galena (Pb),
sulphides tetrahedrite (Cu, Ag, Sb), pyrite (low T),
pyrrhotite (high T)
gangue calcite, fluorite, quartz, dolomite/ankerite/siderite,
baryte/celestite
supergene zone bornite (Cu, Fe), chalcocite (Cu), silver, copper
oxidised zone malachite (Cu), azurite (Cu), cerussite (Pb), anglesite
(Pb), smithsonite (Zn), goethite
As-Sb-Hg ORES low-T orpiment (As), realgar (As), stibnite (Sb), cinnabar (Hg)
hydrothermal
Sn-Mo-W-Bi ORES high-T arsenopyrite (Fe, As), cassiterite (Sn), wolframite (W),
hydrothermal bismuthinite (Bi), molybdenite (Mo), topaz, tourmaline
Ni-Pt ORES mafic magmatic chalcopyrite (Cu), pentlandite (Ni), pyrrhotite (Fe),
platinum (Pt), chromite/magnetite (Cr)
FELSIC PEGMATITES quartz, albite, K-feldspar, muscovite (+ Li micas), beryl
(Be), topaz (F), tourmaline (Li, B), apatite (P, F), zircon
(Zr), monazite (REE, P), ores of lithophiles like Li, Y,
Nb, Ta, W, Th, U
REGOLITH quartz, kaolinite, illite, smectite, hematite, goethite,
calcite, ‘bauxite’ (=diaspore, gibbsite etc)
EVAPORITES halite, gypsum, sylvite, borax
META-ULTRAMAFICS serpentine, talc, anthophyllite, chromite/magnetite,
brucite, magnesite
VESICLE FILLS IN low-T calcite, aragonite, quartz, smectites, zeolites, pectolite,

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 48


BASALTS hydrothermal prehnite, pumpellyite, epidote

Table C5: Summary physical and optical properties of common minerals

his table includes all species that you are likely to see in the course, plus a few more. Comparison
of actual specimens with their descriptions here should help to give you an idea of what the various
terms for cleavage, lustre, habit etc mean, and help you to get to know some of the commonest
minerals quickly. Less common minerals are tyhen easier to identify by comparison with the ones
that you already know. The table is arranged by anion type, in the opposite order to many
textbooks, so that it starts with the very common species quartz and the feldspars.

Note how many of the mineral species listed here fall into family groups (e.g. pyroxenes,
amphiboles, micas, garnets) with distinctive family characteristics.

The table has been split into two parts, respectively for properties that can been seen in hand
specimen properties and those visible in thin section. The ‘habit’ and ‘cleavage’ columns are
duplicated in both.

(i) Hand Specimens

Mineral Anion Type Mohs Colour Lustre Colour Habit Cleavage


Hard- (hand (hand (streak)
ness specimen) specimen)
Quartz Framework 7 various vitreous colourless hexagonal none
Silicate prisms +
rhombohedra
irregular
Plagioclase Framework 6 usually vitreous colourless blades, laths, 2 at almost 90o
(Albite- Silicate white-grey prisms
Anorthite)
Alkali feldspar Framework 6 usually vitreous colourless laths-tablets 2 at 90o
(Albite- Silicate white-grey
Orthoclase/ or pink
Microcline/
Sanidine)
Nepheline Framework 6 grey greasy colourless hexagonal poor-none
(feldspathoid) Silicate prisms
Leucite Framework 6 white-grey greasy colourless trapezohedra none
(feldspathoid) Silicate
Analcime Framework 5.5 white- subvitreous colourless trapezohedra none
(feldspathoid/ Silicate colourless
zeolite)
Cordierite Framework/ 7 blue-grey vitreous colourless pseudo- none
Ring pleochroic! hexagonal
Silicate prisms
Beryl Framework/ 8 various vitreous colourless hexagonal poor basal
Ring prisms
Silicate
Muscovite Layer 3 pale pearly (difficult to plates 1 basal
(mica) Silicate colours, powder)
silvery
Biotite Layer 3 black, green vitreous, (difficult to plates 1 basal
(mica) Silicate “PVC-like” powder)
Phlogopite Layer 3 brown, vitreous, (difficult to plates 1 basal
(mica) Silicate yellow powder)
Clinochlore Layer 3 (dark) green pearly green-grey plates 1 basal
(chlorite) Silicate

Serpentine Layer 2.5-5 various, greasy, silky colourless platy, fibrous, depends on

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 49


group Silicate often green massive habit
Talc Layer 1 white, pale pearly colourless flaky 1 basal
Silicate green

Pyrophyllite Layer 1 white, pale pearly colourless flaky 1 basal


Silicate green

Kaolinite Layer 1-2.5 white dull v. fine-grained 1 basal


Silicate
Prehnite Layer 6-6.5 pale green, vitreous white balls of platy 1 basal
(atypical layer Silicate white xtals, rare
silicate) prisms
Hornblende Double 6 black vitreous grey- green prisms 2 at 54o
(clino-amphibole) Chain
Silicate
Actinolite Double 6 green vitreous grey- green prisms, blades, 2 at 54o
(clino-amphibole) Chain needles
Silicate
Tremolite Double 6 white vitreous colourless prisms, blades, 2 at 54o
(clino-amphibole) Chain needles
Silicate
Glaucophane Double 6 blue-violet vitreous grey prisms-needles 2 at 54o
(clino-amphibole) Chain
Silicate
Anthophyllite Double 6 brown vitreous colourless prisms, needles 2 at 54o
(ortho- Chain
amphibole) Silicate
Augite Single 6.5 black - dark vitreous grey-green equant-prisms 2 at 87o
(clinopyroxene) Chain green
Silicate
Diopside Single 6.5 green-white vitreous colourless prisms 2 at 87o
(clinopyroxene) Chain
Silicate
Omphacite Single 6.5 bright green vitreous colourless prisms 2 at 87o
(clinopyroxene) Chain
Silicate
Jadeite Single 6 green, pale vitreous colourless massive, fibres 2 at 87o
(clinopyroxene) Chain colours
Silicate
Enstatite Single 6 brown vitreous pale brown- equant-tablets 2 at 87o
(orthopyroxene) Chain grey
Silicate

Wollastonite Single 5 white vitreous colourless needles-blades 3 prismatic at


(pyroxenoid) Chain 26o, 70o, 84o
Silicate
Tourmaline Ring 7 various, vitreous greyish striated none
group Silicate often black triangular
(Fe-rich) prisms
Epidote- Pair 6-6.5 green vitreous colourless, prisms 1 prismatic
Clinozoisite Silicate (yellowish) grey, green
Zoisite Pair 6 grey, brown vitreous colourless prisms 1 prismatic
Silicate (pink, blue
rare)
Pumpellyite Pair 6 grey-green vitreous colourless- blades 1 basal +
Silicate green/grey 1 prism
Chloritoid Single 6.5 green-black vitreous- green-grey plates 1 basal + a few
Silicate pearly prismatic
Topaz Single 8 pale colours vitreous colourless rhombic 1 basal
Silicate prisms
Staurolite Single 7.5 brown vitreous colourless stout prisms 1 prismatic
Silicate
Kyanite Single 4-7 blotchy blue vitreous- colourless blades 1 perfect,
Silicate pearly 1 good
Andalusite Single 7.5 grey, pink vitreous colourless prisms, blobs 3 prismatic
Silicate

Sillimanite Single 7 white, grey vitreous, colourless ‘square’ prisms, 1 prismatic


Silicate silky fibres

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 50


Almandine Single 7.5 red-brown vitreous colourless rhombic none
(Fe2+-Al garnet) Silicate dodecahedra,
Pyrope trapezohedra
(Mg-Al garnet)
Grossular Single 7 green, vitreous colourless rhombic none
(Ca-Al garnet) Silicate beige, white dodecahedra,
trapezohedra
Forsterite Single 7 green vitreous colourless equant poor-none
(olivine) Silicate
Zircon Single 7.5 brown, adamantine colourless square prisms, not distinct
Silicate colourless dipyramids
Titanite Single 5 brown adamantine colourless wedge-shaped 2 inclined
Silicate indistinct
Apatite (group) Phosphate 5 various, vitreous colourless hexagonal poor-none
often green prisms
Monazite Phosphate 5 red-brown sub- colourless prism-equant,
adamantine wedge-shaped
Gypsum Sulphate 2 various pale vitreous- colourless prisms, blades, 1 perfect perp to
pearly tablets 2 at 66o
Baryte Sulphate 3.5 various pale vitreous colourless prisms, blades, 1 perfect perp to
tablets 3 less good
Calcite Carbonate 3 various vitreous colourless prisms, rhombs, 3 at 102o
scalenohedra
Dolomite Carbonate 3.5 various vitreous colourless prisms, rhombs, 3 at 102o
scalenohedra
Siderite Carbonate 4-4.5 brown vitreous paler brown rhombs 3 at 102o
Rhodochrosite Carbonate 3.5-4 pink vitreous white-pink rhombs, 3 at 102o
stalactites,
banded massive
Smithsonite Carbonate 3.5-4 various vitreous white rhombs, 3 at 102o
botryoidal
Magnesite Carbonate 3.5-4 white vitreous white rhombs 3 at 102o
Aragonite Carbonate 4 various vitreous colourless prisms, blades prismatic, not
distinct
Cerussite Carbonate 3.5 white adamantine colourless prisms, blades prismatic, not
distinct
Malachite Carbonate 4 green sub- green blades, needles, 1 perfect,
adamantine botryoidal 1 less so
Azurite Carbonate 4 deep blue sub- blue prisms, blades prismatic, not
adamantine distinct
Fluorite Halide 4 various subvitreous colourless cubes, octahedra 4 octahedral
(Fluoride)
Halite Halide 2.5 various pale vitreous colourless cubes, often 3 cubic
(Chloride) skeletal
Rutile Oxide 6.5 brown, red, adamantine- pale brown square prisms 4 prismatic
black submetallic
Cassiterite Oxide 7 brown, adamantine beige square prisms, poor-none
yellow, dipyramids
black
Wolframite Oxide 4.5 black submetallic- brown-black blades 1 perfect
(to red-brn) adamantine
Spinel Oxide 8 various, vitreous colourless- octahedra none
often green- pale
black
Magnetite Oxide 6.5 black submetallic black octahedra none
Chromite Oxide 5.5 black greasy- brown-black octahedra none
submetallic
Corundum Oxide 9 various vitreous scalenohedra basal and
rhomb. parting
Hematite Oxide 6 grey-red metallic- dark red prisms, tablets, basal and
dull botryoidal rhomb. parting
Ilmenite Oxide 6 black submetallic brown-black tablets basal and
rhomb. parting
Goethite Oxide- 5.5 dk. brown - submetallic yellow- botryoidal- 1 good,
Hydroxide mustard to dull brown fibrous 1 less so
yellow
Pyrite Sulphide 6 pale brassy metallic grey-black cubes, none
yellow pentagonal

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 51


dodecahedra
Chalcopyrite Sulphide 4 deep brassy metallic green-black pseudo- poor-none
yellow tetrahedra
Pyrrhotite Sulphide 4 bronze metallic black hexagonal basal parting
brown tablets
Sphalerite Sulphide 4 brown, red, adamantine- pale brown tetrahedra 6 rhombic
black resinous dodec.
Arsenopyrite Sulphide 6 silvery metallic black wedge-shaped poor-none
Bornite Sulphide 3 bronze, metallic brown-black massive, rare none
purple cubes
tarnish
Pentlandite Sulphide 4 brassy metallic black granular octahedral
yellow-pale massive parting
bronze
Galena Sulphide 2.5 grey metallic black cubes, octahedra 3 cubic
Cinnabar Sulphide 2.5 red adamantine red massive, 3 prismatic
rhombohedra

Bismuthinite Sulphide 2 grey metallic black blades, prisms 1 perfect


prismatic
Stibnite Sulphide 2 grey metallic black blades, prisms 1 perfect
prismatic
Orpiment Sulphide 2 yellow adamantine- yellow plates, prisms 1 perfect
pearly
Realgar Sulphide 1.5 red adamantine orange- equant none
yellow
Molybdenite Sulphide 1.5 blue-grey metallic bluish grey plates 1 basal
Graphite Native 1.5 grey submetallic grey plates 1 basal
element
Diamond Native 10 various adamantine- colourless octahedra 4 octahedral
element greasy
Sulphur Native 2 yellow adamantine- pale yellow bipyramids many poor
element resinous
Copper Native 3 copper red, metallic coppery leaves, malleable
element brown dendrites,
tarnish nuggets
Silver Native 3 silvery, metallic silvery leaves, malleable
element brown-black dendrites,
tarnish nuggets
Gold Native 3 deep yellow metallic deep yellow leaves, malleable
element dendrites,
nuggets
Platinum Native 4.5 steely grey metallic steely grey nuggets, cubes malleable
element

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 52


(ii) Thin Section

Mineral Habit Colour Relief Maximum Extinction Twins Cleavage Length


(thin (thin Birefring- slow/
section) section) ence fast
Quartz hexagonal colourless low low 1st straight not visible none slow
prisms + order
rhombohedra
irregular
Plagioclase blades, laths, colourless low low 1st inclined repeated 2 at either
(Albite- prisms order stripes almost 90o
Anorthite)
Alkali feldspar laths-tablets colourless low low 1st inclined simple; 2 at 90o either
(Albite- order X-hatch in
Orthoclase/ microcline
Microcline/ and
Sanidine) anortho-
clase
Nepheline hexagonal colourless low low 1st straight poor-none fast
(feldspathoid) prisms order
Leucite trapezohedra colourless low low 1st n/a fine none n/a
(feldspathoid) negative order X-hatch/
lamellar
Analcime trapezohedra colourless low low 1st n/a fine none n/a
(feldspathoid/ negative order X-hatch/
zeolite) lamellar
Cordierite pseudo- colourless, low 1st order straight sector none either
hexagonal yellow twins
prisms pleo. halos
Beryl hexagonal colourless low 1st order straight poor basal fast
prisms
Muscovite plates colourless medium 3rd-order nearly 1 basal slow*
(mica) straight
Biotite plates green, medium 3rd-4th nearly 1 basal slow*
(mica) yel-brn order straight
pleochroic
Clinochlore plates green-blue low 1st order nearly 1 basal slow*
(chlorite) pleochroic (anomalous straight
colours)

Serpentine platy, various low 1st order usually depends


group fibrous, undulating on habit
massive
Talc flaky colourless low 3rd order nearly 1 basal slow*
straight
Kaolinite v. fine- 1st order nearly 1 basal
grained straight
Hornblende prisms green- medium 1st-2nd inclined simple 2 at 54o slow
(clino- brown order
amphibole) pleochroic
Actinolite prisms, pale green medium 1st-2nd inclined 2 at 54o slow
(clino- blades, pleochroic order
amphibole) needles
Tremolite prisms, colourless medium 1st-2nd inclined 2 at 54o slow
(clino- blades, order
amphibole) needles
Glaucophane prisms- violet- medium 1st-2nd inclined 2 at 54o slow,
(clino- needles blue- order but fast
amphibole) colourless if high
pleochroic Fe!
Anthophyllite prisms, colourless medium 1st-2nd straight 2 at 54o slow
(ortho- needles order
amphibole)
Augite equant- pale beige, medium 2nd order inclined simple 2 at 87o either
(clinopyroxene) prisms purple

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 53


Diopside prisms colourless medium 2nd order inclined simple 2 at 87o either
(clinopyroxene)
Omphacite prisms pale green medium 2nd order inclined simple 2 at 87o either
(clinopyroxene)
Jadeite massive, colourless medium 1st order inclined 2 at 87o either
(clinopyroxene) fibres
Enstatite equant- pink-green medium 1st order straight none! 2 at 87o slow
(orthopyroxene) tablets pleochroic
Wollastonite needles- colourless medium 1st order inclined 3 either
(pyroxenoid) blades prismatic
at 26o,
70o, 84o
Tourmaline striated various medium 2nd order straight none fast
group triangular pleochroic
prisms
Epidote prisms green- high 2nd-3rd inclined 1 either
yellow order prismatic
pleochroic
Clinozoisite prisms colourless high 1st order inclined 1 either
anomalous prismatic
Zoisite prisms colourless high 1st order straight 1 fast
anomalous prismatic
Pumpellyite blades blue-green- medium- 1st-2nd inclined 1 basal + slow
yellow high order 1 prism
pleochroic
Chloritoid plates blue-green- high 1st order nearly lamellar 1 basal + fast*
yellow anomalous straight a few
pleochroic prismatic
Topaz rhombic colourless medium 1st order straight 1 basal slow
prisms
Staurolite stout prisms yellow high 1st order straight X-twins 1 slow
pleochroic prismatic
Kyanite blades colourless high 1st order inclined 1 perfect, slow
1 good
Andalusite prisms, blobs colourless medium 1st order straight 3 fast
prismatic
Sillimanite ‘square’ colourless medium 2nd order straight 1 slow
prisms, prismatic
fibres
Almandine rhombic pink high none n/a none n/a
(Fe2+-Al garnet) dodecahedra,
Pyrope trapezohedra
(Mg-Al garnet)
Grossular rhombic colourless high none n/a none n/a
(Ca-Al garnet) dodecahedra,
trapezohedra
Forsterite equant colourless medium 2nd-3rd straight poor-none
(olivine) order
Zircon square colourless v. high 3rd-4th straight not
prisms, order distinct
dipyramids
Titanite wedge- pale brown v. high v. high inclined 2 inclined
shaped order indistinct
Apatite (group) hexagonal colourless medium 1st order straight poor-none fast
prisms
Monazite prism- colourless- high 3rd-4th nearly
equant, pale brown order straight
wedge-
shaped
Gypsum prisms, colourless low 1st order inclined swallow 1 perfect
blades, tails ⊥ 2 at 66o
tablets
Baryte prisms, colourless medium 1st order straight 1 perfect
blades, ⊥ 3 less
tablets good
Calcite prisms, colourless low-high v. high straight lamellae || 3 at 75o fast
rhombs, (twinkling order (bisects edges,
scalenohedra ) cleavage) long
diagonal

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 54


of cl.
rhombs
Dolomite prisms, colourless low-high v. high straight lamellae || 3 at 75o fast
rhombs, (twinkling order (bisects both diags
scalenohedra ) cleavage) of
cleavage
rhombs
Aragonite prisms, colourless low-high v. high straight pseudo- prismatic, fast
blades (twinkling order hexag. not
) trillings distinct
Cerussite prisms, colourless high v. high straight pseudo- prismatic,
blades order hexag. & not
V-twins distinct
Malachite blades, deep green medium- inclined 1 perfect,
needles, pleochroic v.high 1 less so
botryoidal
Azurite prisms, deep blue high inclined prismatic,
blades pleochroic not
distinct
Fluorite cubes, colourless medium none n/a 4 n/a
octahedra negative! octahedral
Halite cubes, often colourless v. low none n/a 3 cubic n/a
skeletal
Rutile square brown v. high v. high straight knee twins 4
prisms prismatic
Cassiterite square colourless- v. high v. high straight knee twins poor-none
prisms, pale
dipyramids
Wolframite blades opaque v. high n/a n/a 1 perfect
(red-brn)
Spinel octahedra often green high none n/a simple none n/a
‘spinel
law’
Magnetite octahedra opaque n/a n/a n/a none
Chromite octahedra opaque n/a n/a n/a none
Corundum scalenohedra colourless high 1st order straight basal and fast
rhomb.
parting
Hematite prisms, red if thin v. high v. high basal and
tablets, rhomb.
botryoidal parting
Ilmenite tablets opaque n/a n/a n/a basal and
rhomb.
parting
Goethite botryoidal- orange v. high v. high 1 good,
fibrous if thin 1 less so
Pyrite cubes, opaque n/a n/a n/a none
pentagonal
dodecahedra
Chalcopyrite pseudo- opaque n/a n/a n/a poor-none
tetrahedra
Pyrrhotite hexagonal opaque n/a n/a n/a basal
tablets parting
Sphalerite tetrahedra brown- v. high none n/a 6 rhombic
opaque dodec.
Arsenopyrite wedge- opaque n/a n/a n/a poor-none
shaped
Bornite massive, rare opaque n/a n/a n/a none
cubes
Pentlandite granular opaque n/a n/a n/a octahedral
massive parting
Galena cubes, opaque n/a n/a n/a 3 cubic
octahedra
Cinnabar massive, deep red v.high v.high n/a simple 3
rhombohedra prismatic
Bismuthinite blades, opaque n/a n/a n/a 1 perfect
prisms prismatic
Stibnite blades, opaque n/a n/a n/a deformatn 1 perfect

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 55


prisms . prismatic
kinks
Orpiment plates, yellow v.high v.high n/a 1 perfect
prisms
Realgar equant red v.high v.high n/a none
Molybdenite plates opaque n/a n/a n/a 1 basal
Graphite plates opaque n/a n/a n/a 1 basal
Diamond octahedra colourless v. high none n/a simple 4
‘spinel octahedral
law’
Sulphur bipyramids yellow v. high v. high straight many poor

*Length slow/fast information is asterisked if the ‘length’ referred to is actually the width of a platy
crystal. Be very careful about checking the 3-D habit of crystals when using this information.

(iii) Unusual tenacity

Most minerals are brittle. Some that show other properties are:
Flexible and elastic: normal micas such as biotite and muscovite.
Flexible but not elastic: chlorite, talc, pyrophyllite, graphite, molybdenite, orpiment
Sectile: stibnite

(iv) Unusual density

Quartz, feldspars and calcite all have densities near 2.6-2.7 g/cm3, so this is a very typical figure for
minerals of the continental crust. Some species that are significantly denser, which is very
noticeable if they are transparent or nonmetallic. Others, usually hydrous minerals, are distinctly
light. In more detail:

Density:
2.0-2.5: Tridymite/Cristobalite, Leucite, Sodalite, Zeolites, Brucite, Gibbsite, Gypsum, Halite,
Graphite, Sulfur

2.5-3.0: Beryl, Cordierite, Talc, Chlorite, Serpentine, Kaolinite, Muscovite, Phlogopite,


Nepheline, Scapolite, Anhydrite, Calcite, Magnesite, Dolomite, Quartz, Feldspars

3.0-3.25: Sillimanite, Andalusite, Pumpellyite, Tourmaline, Fe-rich biotite, Apatite, Fluorite

3.25-3.5: Forsterite, Epidote, Enstatite, Diopside-Hedenbergite-Augite, Jadeite, Diaspore,


Orpiment, Realgar

3.5-4.5: Fayalite, Ferrosilite, Titanite, Garnet, Topaz, Kyanite, Staurolite, Chloritoid, Periclase,
Corundum, Spinel, Goethite, Chalcopyrite, Sphalerite, Celestine, Siderite, Rhodochrosite,
Strontianite, Stishovite, Diamond

4.5-5.5: Zircon, Hematite, Ilmenite, Rutile, Magnetite, Chromite, Pyrite, Pyrrhotite, Bornite,
Molybdenite, Stibnite, Baryte, Monazite

5.5-9: Cassiterite, Galena, Bismuthinite, Arsenopyrite, Cinnabar, Nickeline, Cerussite, Native Iron,
Native Copper

> 9: Gold, Platinum, Silver, Mercury

Andrew G. Christy - Mineral Identification Handbook 4th ed (2010) 56

You might also like